Dear reader of ADxS.org, please excuse the disruption.

ADxS.org needs about $63500 in 2024. In 2023 we received donations of about $ 32200. Unfortunately, 99.8% of our readers do not donate. If everyone who reads this request makes a small contribution, our fundraising campaign for 2024 would be over after a few days. This donation request is displayed 23,000 times a week, but only 75 people donate. If you find ADxS.org useful, please take a minute and support ADxS.org with your donation. Thank you!

Since 01.06.2021 ADxS.org is supported by the non-profit ADxS e.V..

$8975 of $63500 - as of 2024-02-29
14%
Header Image
MPH Part 1: Active ingredients, effect, responding

Sitemap

MPH Part 1: Active ingredients, effect, responding

1. Active ingredient

Methylphenidate (MPH):1

  • belongs to the phenylethylamine class.
  • Chemical name: Methyl-2-phenyl-2-(piperidin-2-yl)acetate
  • Molecular formula: C14H19NO2
  • Mass: 233.31 g/mol
  • has four configuration isomers
  • As methylphenidate has two asymmetric carbon atoms, there are four different forms of this drug2
    • (+)-Erythromethylphenidate
    • (-)-Erythromethylphenidate
    • (+)-threomethylphenidate
    • (-)-threomethylphenidate.

MPH is classified worldwide as a narcotic because it can be abused as a drug when taken in extreme doses and forms. However, when taken in medication doses and in medication form (oral/patch), MPH has no potential for intoxication or dependence. The actual risk of abuse of methylphenidate as a drug appears to be considerably overestimated. A study that claimed frequent misuse of MPH as a drug merely referred to sources in which methylphenidate is not even mentioned3

With (especially older) studies on the effect of MPH, it must always be taken into account that these4

  • MPH usually used in significantly higher doses than for ADHD medication
    • Although rodents require higher doses than humans, doses are often used that correspond to drug intake and not to medication intake
  • use unretarded / not prolonged-acting MPH via prodrug
  • frequently inject MPH, which can result in much faster metabolism depending on the type of injection
    • IP injection should be similar to oral intake
  • these 3 factors multiply in their effect

There is no doubt that MPH in medication form has a different effect than MPH in drug form.

Helpful details for assessing the validity of MPH animal studies

Recording forms:

Intraperitoneal injections (IP, injections into the abdominal cavity)

  • IP in rodents is performed in the lower right quadrant of the abdomen below the midline. This video shows how such an IP injection is performed (German). IP injections should largely correspond to oral ingestion. Absorption is much slower intraperitoneally than with intravenous injections. IP allows for more efficient absorption of MPH in the mesenteric vessels that enter the portal vein and pass through the liver, allowing given drug to undergo hepatic metabolism before reaching the systemic circulation. In addition, a small amount of the intraperitoneally injected material can pass through small lacunae directly through the diaphragm into the thoracic lymph5

  • MPH administration via IP injection increases the DA concentration in the brain faster and much longer lasting than oral administration.6

  • IP injection of drugs in experimental studies with rodents is acceptable for pharmacological and proof-of-concept studies to investigate the effect(s) of the target effect. It is unsuitable for studies on the properties of a drug formulation and/or its pharmacokinetics for clinical implementation.7

  • The risks of IP injections are:

    • Injuries due to incorrect injection
    • Change in effect if injection too close to the surface (subcutaneous instead of intraperitoneal)
  • Subcutaneous injections have the following effects

    • MPH half-life shortened
    • Peak release of dopamine in the brain increased8

Oral gavage
Oral gavage better mimics oral uptake and human MPH metabolism. Oral administration uses a suitable tube or administration needle that is inserted into the animal’s mouth and esophagus. However, the manual fixation required for oral gavage causes such severe stress in both rats and mice that it increases plasma corticosterone levels9

Further recording channels:5

  • enteral (through the digestive tract)
  • parenteral (outside the digestive tract)
    • avoid the first-pass effect of hepatic metabolism, which often occurs with oral administration
      • therefore generally generate a higher bioavailability
    • avoid unpredictable effects associated with enteral absorption processes
  • orally (in the mouth)
  • directly into the stomach (gastric tube)
  • intravenously (into a blood vessel)
  • epicutaneous (on the skin)
  • intradermal (into the skin)
  • subcutaneously (under the skin)
  • transdermal (through the skin, e.g. plaster)
  • intramuscular (into the muscle)
  • transcorneal (on the eye)
  • intraocular (or into the eye)
  • intracerebral (into the brain)
  • epidural (in the space surrounding the dura mater)
  • intrathecal (in the space surrounding the distal spinal cord);
  • intraosseous (into the bone marrow cavity)
  • intranasal (sprayed into the nose for absorption via the nasal mucous membranes or the lungs)
  • intratracheal (into the lungs by direct tracheal instillation or inhalation)

The dosage does not differ between mice and rats on the basis of the species, but solely on the basis of the size of the respective animal.
MPH doses act according to an inverted-U principle10

  • In rodents (mice such as rats):
    • Doses of 1 mg/kg MPH or less:
      • no effects on movement activity due to subcutaneous injection11
    • Doses of 2 to 5 mg/kg MPH appear to correspond to a drug dose in humans12
      • by subcutaneous injection: increase in movement activity
    • Doses of 10 to 20 mg/kg MPH are said to be more like drug intake13
      • by subcutaneous injection: increase in movement activity
        • Maximum at 10 mg / kg11
    • Doses of 40 mg / kg MPH
      • by subcutaneous injection: more concentrated, stereotyped activity that reduced the amount of forward movement11

1.1. Methylphenidate as racemate

Methylphenidate is usually offered clinically as a racemate (mixture) of 50 % L-methylphenidate and 50 % D-methylphenidate (levorotatory and dextrorotatory isomers), which is ten to a hundred times more effective than the (+/-) erythroform.2
While a significant proportion of D-MPH enters the CNS via the blood-brain barrier, L-MPH is not absorbed into the CNS.14

Brand names include Ritalin, Medikinet, Equasym, Concerta, Kinecteen, Daytrana (skin plaster). It is also offered as a generic. See below for more information.

Dosage from 2.5 mg to an average of 15 mg per single dose during the day every 2.5 to 3.5 hours (unretarded form).15

1.2. Dexmethylphenidate (D-MPH)

Dexmethylphenidate is the pure form of the dextrorotatory isomers and is pharmacologically active.

Brand name: Focalin (Switzerland and USA only)

It is 3 times more effective than racemic (mixture of dextrorotatory (D-MPH) and levorotatory (L-MPH) methylphenidate.16 The higher efficacy of D-MPH compared to L-MPH concerns the dopamine transporter as well as the noradrenaline transporter binding.
It is therefore recommended to halve the dosage of D-MPH compared to racemate methylphenidate and to limit it to a maximum of 20 mg/day in children and adults.
D-MPH is also available as a prolonged-release preparation (Focalin XR).

1.3. Levomethylphenidate (L-MPH)

L-MPH is the pure form of the levorotatory isomers and is pharmacologically almost ineffective.
L-MPH is predominantly metabolized in the intestinal mucosa. The remaining L-MPH is metabolized by the liver during the first pass. L-MPH is almost completely metabolized within 10 minutes of ingestion17
The small, unmetabolized residue of L-MPH competes with the dextrorotatory and pharmacologically active isomer D-MPH for the binding sites in the brain and reduces its effectiveness.17
According to another source, the left-turning MPH isomer L-MPH (unlike D-MPH) is not absorbed into the CNS.18

1.4. Serdex methylphenidate (SDX)

Serdexmethylphenidate (SDX) is a prodrug of dexmethylphenidate (d-MPH) and is currently only approved in the USA (Azstarys®, KP415 with 70 % SDX and 30 % MPH).19

The advantage of prodrugs is a prolonged effect, as the conversion of product into active ingredient delays the release. See also lisdexamfetamine, which is first converted to dextroamphetamine (Elvanse).

2. Mode of action of methylphenidate

The first computer models now exist that can seriously simulate the effect of ADHD drugs. A computer model for the simulation of type 1 diabetes has already been approved by the FDA as a replacement for preclinical animal studies20
A model comparing MPH and AMP in children and adults with ADHD takes into account the effect on 99 proteins involved in ADHD21

2.1. Effect on neurotransmitters

MPH changes the neurotransmitter levels in the brain.

2.1.1. MPH and dopamine

2.1.1.1. MPH increases extracellular and possibly decreases phasic dopamine

Many sources only report a general effect on dopamine and noradrenaline.
The blanket statement that MPH increases dopamine is ultimately not helpful, as a distinction must be made between tonic and phasic dopamine release and extracellular dopamine levels, as well as between the effect in different regions of the brain.

  • Increase in dopamine and noradrenaline in the PFC222324
  • MPH doses whose plasma levels are in the range of therapeutic use for ADHD, such as2526
    • 1.0 mg/kg in rats
      • Increase extracellular noradrenaline in the hippocampus by 100
      • Do not significantly alter dopamine in the nucleus accumbens (striatum)
    • 2.5 mg/kg in rats
      • Increase extracellular noradrenaline in the hippocampus by 150-175 %
      • Do not significantly alter dopamine in the nucleus accumbens (striatum).
  • In healthy adult rats, MPH increases
    • Dopamine in the PFC, striatum and nucleus accumbens and27
      • High doses: DA increased by release from noradrenergic cells28
      • Low dosage: DA not increased28
      • Chronic administration of 1 mg/kg increased extracellular dopamine and noradrenaline in the PFC29
    • Noradrenaline in the PFC, but not in the striatum or nucleus accumbens.27
    • Dopamine and noradrenaline in the PFC, but hardly in other areas of the brain30
  • In healthy monkeys, MPH:31
    • Striatum: low and high doses: consistent increase in dopamine
    • PFC: increase in dopamine only at high doses
      In rats, 1 mg/kg MPH as a single dose as chronic (21 days) caused
    • Striatum: no change in extracellular levels of noradrenaline, dopamine or serotonin29

In summary, MPH appears to affect only the PFC and hippocampus in rats, whereas in primates it also affects the striatum.

Tonic dopamine mediates the regulatory (inhibitory) control of the PFC on the ventral striatum, thus inhibiting the (phasic) activity of the striatum. In response to unexpected positive reward stimuli, the striatum fires phasically in a dopaminergic manner and activates dopaminergic postsynaptic receptors. Tonic control is therefore inhibitory and modulates excitatory phasic firing in response to unexpectedly positive reward stimuli.32

Stimulants33

  • increases the tonic rate of fire
  • shortens the frequency of postsynaptic multisecond oscillations in the basal ganglia from around 30 seconds to 5 to 10 seconds. Already 1.0 mg/kg MPH or 0.2 mg/kg D-Amp showed an effect.
  • however, 0.2 mg/kg D-Amp showed only a slight presynaptic effect. D-Amp had a stronger postsynaptic effect.

MPH increases extracellular dopamine and decreases phasic dopamine in the striatum.
This is consistent with Grace’s hypothesis that ADHD is characterized by decreased tonic and increased phasic dopamine. This model is likely to apply at least to most cases of ADHD. More on this under Dopamine release (tonic, phasic) and coding

2.1.1.1.1. MPH increases extracellular dopamine

MPH increases extracellular dopamine 34 35 3628

The statement “MPH increases tonic dopamine” would not be correct because tonic firing itself is not increased by MPH. The inhibition of reuptake only leads to an increase in the extracellular dopamine level, not to a change in tonic firing.

The following reasons for the increase in extracellular dopamine by MPH were identified

  • an inhibition of dopamine reuptake34
  • increased dopamine efflux due to reversal of the dopamine reuptake transporter in the striatum28

MPH increased response strength and reward expectancy-related BOLD signaling in the ventral striatum during a gambling task,36 suggesting that striatal tonic dopamine levels represent an average reward expectancy signal that modulates the phasic dopaminergic response to reward.

The DAT blocker nomifensine enhanced the phasic dopaminergic signal34, so dopamine reuptake inhibition per se cannot be the only reason for the reduction in tonic dopamine by MPH

2.1.1.1.2. MPH can reduce phasic dopamine

Several studies found a reduction in phasic dopamine by MPH,3437 36 but some also found no change3528 or an increase. This could depend on the circumstances (e.g. dosage, DAT sensitivity).

The significant reduction in the phasic (synaptic) release of dopamine is probably due to a decrease in synapsin phosphorylation34
The inhibition of phasic dopamine is not a consequence of presynaptic D2 autoreceptor activation, since this34
- a. also in the presence of the D2 antagonist sulpiride and
- b. also occurs in DAT-CI striatal slices in which activation of D2 autoreceptors cannot occur due to an increased extracellular DA level
The inhibition of phasic dopamine could possibly be due to a reduction in synaptic vesicle neurotransmitter, as with cocaine.34 Cocaine, as a lipophilic weak base, could collapse the vesicular pH gradient, similar to the “weak base effects” reported for amphetamines, or act on vesicles. There is evidence that synaptic vesicles can be quite “leaky” and constantly leak dopamine into the cytosol. This impression is supported by the action of reserpine, which can rapidly deplete synaptic vesicles of dopamine.

One study found that MPH did not alter phasic dopamine, suggesting an inhibitory feedback mechanism via D2 autoreceptors, because a D2 antagonist given in parallel with MPH caused MPH to also increase phasic dopamine35

Dopamine reuptake inhibitors such as MPH should generally lead to increased phasic dopamine in the dorsolateral striatum37
Acute MPH administration increased the firing activity of PFC pyramidal neurons in rats and potentiated NMDA-induced neurotransmission28

2.1.1.1.3. Stimulants alter dopaminergic firing rate

Stimulants33

  • increase the dopaminergic firing rate in the globus pallidus in a dose-dependent manner
    • MPH by up to 100 %
    • D-AMP by up to 50 %
  • reduce the dopaminergic firing rate in the substantia nigra pars compacta
    • MPH by up to 100 %
    • D-AMP by up to 100 %
2.1.1.2. MPH binds to DAT (reuptake inhibition)

Methylphenidate as a dopamine reuptake inhibitor increases the dopamine level in the synaptic cleft.22 MPH inhibits the dopamine transporter and the noradrenaline transporter38 From this it could be concluded that the site of action of MPH is where there is a dopamine deficiency. In the mesocortical model of ADHD, this would be the PFC. However, SPECT and PET studies clearly show that MPH primarily increases dopamine activity in the striatum, which argues against the PFC as the (sole) site of action (which correlates with the low DAT count in the PFC and the high DAT count in the striatum). Since, according to the mesocortical model of ADHD, dopamine activity in the ventral striatum is excessive, MPH, if it has an increasing effect there, should worsen rather than improve symptoms. Low-dose stimulants such as MPH can inhibit phasic dopamine release by enhancing inhibitory tonic control.39 However, in an fMRI study, children with ADHD without medication showed increased frontal and decreased striatal activation, arguing against the mesocortical deficiency theory. MPH increased frontal blood flow in children with and without ADHD, but only increased striatal blood flow in children with ADHD. It is therefore unclear whether the observed frontal deficits in ADHD reflect a central dysfunction in the PFC or a lack of input from other dopaminergic systems. Since almost all mental disorders show a certain degree of frontal dysfunction, it is unclear whether the aetiological deficits in ADHD have other causes.32

  • MPH binds to the dopamine transporters whose density is highest in the striatum. The binding of MPH in the cerebellum and hippocampus is less than a tenth of this.40
  • MPH does not bind to dopamine receptors, but only to DAT and NET.4142 (In contrast: MPH is said to activate postsynaptic D1 receptors.22 )
    • D-MPH binds most strongly, namely to
      • DAT with IC50 = 23 nM, Ki = 161 nM;
      • NET with IC50 = 39 nM, Ki = 206 nM
    • D/l-MPH racemate binds somewhat weaker, namely to41
      • DAT with IC50 = 20 nM, Ki = 121 nM
      • NET with IC15 = 51 nM, Ki = 788 nM
    • L-MPH binds the weakest, namely to41
      • DAT with IC50 = 1600 nM, Ki = 2250 nM
      • NET with IC50 > 104 nM, Ki > 104 nM
  • MPH drug doses cause an MPH plasma concentration of around 20 to 30 nM, which is sufficient to occupy a significant proportion of the dopamine transporters. This effect coincides with that of D-AMP.39
  • MPH responders showed an increased DAT count in the striatum, non-responders a reduced DAT count.43
  • MPH increased the number of dopamine transporters.44
  • Different:
    • Lower increase in dopamine and noradrenaline in the striatum22
    • No increase in dopamine in the striatum by MPH in DAT(-/-) mice, but in DAT(+/-) and DAT(+/+) mice45 Apparently, the degree of expression/sensitivity of the DAT is decisive for the positive or negative modulation of phasic dopamine release.46
2.1.1.3. MPH increases dopamine release through DAT (efflux)

While it was previously assumed that methylphenidate does not cause dopamine release, more recent opinions assume that it does

  • a dopamine efflux from the DAT
  • dopamine release from vesicles at very high doses.

The view that MPH is a pure DA reuptake inhibitor4748 is outdated.

  • Also release of dopamine from vesicles (here: reserpine-sensitive granules)4950
    • Should only occur at very high doses of more than 80 mg / day5152
  • In contrast to amphetamine, methylphenidate is not considered a substrate for transport into the cytoplasm, which is why it causes at most a slight presynaptic release of dopamine.53
  • Dopamine efflux through DAT reversal in the PFC
    • In the striatum in rats at 4 mg/kg (measured ex vivo)28
    • MPH thereby increases the extracellular dopamine level
    • Efflux from dopamine and noradrenaline terminals
    • By vesicular dopamine release and by sodium-dependent mechanisms
  • Increase the firing rate of PFC pyramidal neurons
    • At chronic intake in PFC in rats at 4 mg/kg (measured in vivo)28
    • measured by extracellular electrophysiological single-cell recordings
    • Reactions to locally applied NMDA unchanged
  • Desensitization to both dopamine and MPH in striatal regions
    • With chronic ingestion in the striatum in rats at 4 mg/kg (measured in vivo)28
    • reduced efficacy of extracellular dopamine in modulating NMDA-induced firing activities of medial spinous process neurons in the striatum
    • lower MPH-induced dopamine outflow
    • is consistent with the empirical experience that a one-time adjustment of the MPH dose is required after a few weeks in the case of single dosing
2.1.1.4. MPH acts on DA via D2 autoreceptors

MPH causes the disinhibition (removal of inhibition) of presynaptic D2 autoreceptors54
Nevertheless, the effect of stimulants in different doses on D2 autoreceptors is not greater than on the postsynaptic heteroreceptors. Stimulants appear to have little effect on the dopamine system above autoreceptors.33 * In people with a high number of D2 receptors, MPH increases metabolism in frontal and temporal brain areas (including the striatum), whereas in healthy people with a low number of D2 receptors, MPH decreases metabolism. Metabolism was consistently increased in the cerebellum.55
* This corresponds to a normalization of the D2 receptor binding.56

  • Methylphenidate normalizes increased dopamine transporter densities in ADHD-HI rats more than in ADHD-I rats57
2.1.1.5. MPH increases DA via VMAT2

MPH influences the redistribution of the vesicular monoamine transporter-2 (VMAT-2; Solute Carrier Family 18 Member 2 - SLC18A2). VMAT2 is involved in the sequestration of cytoplasmic dopamine and noradrenaline and is therefore an important regulator of neurotransmission. MPH does not affect the total amount of VMAT-2 in presynaptic terminals, but only VMAT-2 transport.5859
MPH has the following effects in monoaminergic neurons (but not in cholinergic, GABA-ergic or glutamatergic neurons)41

  • Decrease in VMAT-2 immunoreactivity in the membrane-associated fraction
  • Increase in the cytoplasmic fraction
  • no change in the entire synaptosomal pool

MPH thus protects the dopaminergic system from progressive “wear and tear” by securing a considerable DA reserve pool in the presynaptic vesicles. Therefore, there is only a relatively low risk of neurotoxic / neuropsychiatric side effects in treatment practice with MPH41

According to older reports, MPH has no effect on vesicular monoamine transporters (VMAT).38

2.1.1.6. MPH increases tyrosine hydroxylase

Tyrosine hydroxylase (TH) is the rate-limiting enzyme for the synthesis of dopamine. TH converts tyrosine into the DA precursor L-3,4-dihyroxyphenylalanine (L-DOPA). MPH thus supports dopamine synthesis.
MPH (as well as sport) can induce the expression of TH60 and increase TH levels61
d-MPH from 100 nmol/l significantly increased tyrosine hydroxylase activity in vitro; L-MPH or racemic MPH at the same concentration did not increase TH62
It is unclear whether the increase occurs only peripherally or also in the brain.63 TH gene variants appear to influence the response of MPH.63

2.1.1.7. MPH acts via L-DOPA receptors

The L-dopa receptor GPR143 appears to be involved in the acute and chronic effects of MPH.
Although MPH increases dopamine release, it does not affect L-DOPA release from the dorsolateral striatum. Nevertheless, in L-DOPA receptor KO mice (mice with a defect in the Gpr143 L-DOPA receptor gene), L-DOPA release is reduced:64

  • the hyperlocomotion induced by MPH
  • the reward effect
  • the c-fos expression induced by MPH.
2.1.1.8. Dose-dependent effect of MPH

The effect of MPH is dose-dependent. Normal doses of MPH have different effects than high or very high doses of MPH.

  • At low doses, methylphenidate increases dopamine and noradrenaline levels in the PFC, which increases its performance. In contrast, low-dose MPH has hardly any effect on dopamine and noradrenaline levels in other areas of the brain.65 This corresponds to the known increase in cognitive performance of the PFC due to small increases in dopamine and noradrenaline levels during mild stress.
  • At 3 mg/kg MPH, one study found no increase in dopamine or noradrenaline in the striatum of rats66
  • At higher doses of MPH (as well as cocaine), a reduction in dopamine levels in the striatum was reported in a laboratory study in rats. Only lower doses of MPH or cocaine caused increases in dopamine levels. Moreover, this was not the case in all animals.46
2.1.1.9. Duration-dependent effect of MPH

Acute MPH administration increased the firing activity of PFC pyramidal neurons in rats and potentiated NMDA-induced neurotransmission.
Chronic MPH administration (2 x 2 mg/kg/day) showed 28 days after the end of MPH administration on pyramidal neurons of the PFC28

  • long-term increase in firing activity
  • unchanged burst activities
  • unchanged total number of spontaneously discharging neurons
  • unaltered glutamatergic neurotransmission

Long-term administration of MPH or atomoxetine to juvenile Naples High-Excitability (NHE) rats in adulthood67

  • Dopamine in the PFC and striatum reduced
  • Noradrenaline increased in the ventral striatum
  • In juvenile rats5859
    • A single administration of high-dose MPH (2 mg/kg, i.e. approximately twice the maximum treatment dose normally used in humans)
      • A reduction in the number of vesicular monoamine transporters (VMAT2) in the cerebellum
      • No increase in dopamine turnover in the cerebellum (measured by the DA metabolite DOPAC)
      • No change in the protein levels of tyrosine hydroxylase (TH) and the dopamine D1 receptor
      • Unchanged levels of dopamine and homovanillic acid (HVA)
    • Continuous administration of high-dose MPH (2 mg/kg) over 14 days
      • Increased number of vesicular monoamine transporters (VMAT2) in the cerebellum
      • Significantly increased dopamine turnover in the cerebellum (measured with the metabolite DOPAC)58
      • Left the protein levels of tyrosine hydroxylase (TH) and the dopamine D1 receptor unchanged58 - unlike the same authors59
      • Increased the number of DAT59
        • In the left dorsal striatum
      • Did not change the DAT
        • In the right dorsal striatum
        • In the nucleus accumbens (ventral striatum)58
      • Increased the expression of the
        • Noradrenaline transporter (NET)
        • Monoamine transporter 2 (VMAT2)5859
          • In contrast, amphetamine reduces VMAT268
        • Tyrosine hydroxylase
        • Dopamine D1 receptors
        • Stronger in the nucleus accumbens (ventral striatum) than in the dorsal striatum
        • Stronger in the parietal cortex than in the frontal cortex
        • This effect of chronic MPH on increasing DAT, NET and VMAT2 transporters may suggest that the drug could lose some of its acute effect of increasing dopamine and noradrenaline levels in the long term.5944
          This is consistent with our experience that for some users the dosage has to be adjusted (slightly increased) once after six months to a year. However, a general habituation effect is neither reported in studies69 nor in practice.
        • Elevated vanillin mandelic acid in the urine of Wistar rats. This could be avoided by augmentative administration of buspirone.70
          Vanillin mandelic acid is produced during the breakdown of adrenaline and noradrenaline by MAO-A and COMT, so that vanillin mandelic acid is an indicator of the activity of the autonomic nervous system (sympathetic nervous system).
      • A single administration of very high doses of MPH (5 mg/kg, i.e. about 5 to 20 times the usual treatment dose for humans) has the following effects
        • A similar metabolite change in the cerebellum as 2 mg
        • Metabolites in the cerebellum associated with energy expenditure and excitatory neurotransmission, here glutamate, glutamine, N-acetylapartate and inosine, tended to be reduced
        • Furthermore, the levels of some metabolites associated with inhibitory neurotransmission, in this case GABA and glycine, acetate, aspartate and hypoxanthine, were reduced
  • One study found that basal oxytocin levels in children with ADHD were unchanged compared to unaffected children. While oxytocin decreased in untreated ADHD sufferers after interaction with a parent, oxytocin increased in ADHD sufferers treated with MPH in the same way as in unaffected children.71

2.1.2. MPH increases noradrenaline

  • MPH has a noradrenergic effect in the locus coeruleus, which improves arousal, vigilance and attention22
  • The effect of MPH is dose-dependent. Normal doses of MPH have different effects than high or very high doses of MPH.
  • At low doses, methylphenidate increases dopamine and noradrenaline levels in the PFC, which increases its performance. In contrast, low-dose MPH has hardly any effect on dopamine and noradrenaline levels in other areas of the brain.65 This corresponds to the known increase in cognitive performance of the PFC due to small increases in dopamine and noradrenaline levels during mild stress.
2.1.2.1. MPH low dose increased extracellular noradrenaline, but not dopamine

Adolescent rats received 0.75-3.0 mg/kg MPH orally during the dark-active phase of the circadian cycle, which remained below the threshold for locomotor activation. These doses:26

  • increased extracellular norepinephrine in the hippocampus
  • altered dopamine in the nucleus accumbens does not
  • did not alter methamphetamine sensitivity
  • did not cause any habituation effects

10, 20 and 30 mg/kg MPH (far above a medicinal dosage):72

  • stereotypic behavior (a sign of strongly increased extracellular dopamine); 20 mg/kg as strong as 2.5 mg/kg AMP
  • extracellular dopamine increased
  • extracellular noradrenaline increased
  • extracellular serotonin unchanged (unlike AMP, as it is elevated)
2.1.2.2. MPH binds to NET (reuptake inhibition)
  • Noradrenaline reuptake inhibitors2228
  • This also increases extracellular dopamine in the PFC, where there is little DAT but plenty of NET
2.1.2.3. MPH causes noradrenaline efflux in the PFC

2.1. MPH induces dopamine and noradrenaline efflux in the prefrontal cortex
In the prefrontal cortex (PFC), administration of 100 µM methylphenidate (MPH) (Figure 1A) elicited significant ex vivo dopamine release (Bonferroni post-hoc test after significant two-way ANOVA, Supplementary Table S1). This effect was dependent on norepinephrine terminals, as incubation of radiolabeled dopamine (35 nM) in the presence of desipramine significantly attenuated the dopamine efflux induced by 100 µM MPH (Figure 1A, Bonferroni post-hoc test after significant two-way ANOVA). This was further confirmed by assessing radiolabeled norepinephrine efflux (67-83 nM) after MPH exposure in PFC samples. Indeed, MPH at 100 µM induced significant norepinephrine efflux in the PFC (Figure 1B, Tukey’s post hoc test after significant one-way ANOVA). A lower dose of 10 µM MPH did not elicit dopamine efflux under any conditions. These results suggest that MPH can induce both dopamine and norepinephrine efflux in the PFC, an effect that originates from both dopamine and norepinephrine terminals at a dose of 100 µM.

  • Noradrenaline efflux (67-83 nM) from noradrenaline terminals28
    • Measured ex vivo
    • At 100 µM MPH, not at 10 µM MPH
2.1.2.4. MPH increases NE via VMAT2

MPH influences the redistribution of the vesicular monoamine transporter-2 (VMAT-2; Solute Carrier Family 18 Member 2 - SLC18A2). VMAT2 is involved in the sequestration of cytoplasmic dopamine and noradrenaline and is therefore an important regulator of neurotransmission.
MPH does not affect the total amount of VMAT-2 in presynaptic terminals, but only the VMAT-2 transport.
MPH has an effect in monoaminergic neurons (but not in cholinergic, GABA-ergic or glutamatergic neurons41

  • Decrease in VMAT-2 immunoreactivity in the membrane-associated fraction
  • Increase in the cytoplasmic fraction
  • no change in the entire synaptosomal pool
2.1.2.5. MPH binds to noradrenaline receptors

MPH binds directly to noradrenergic receptors.73 MPH binds to41

  • α2A (Ki = 5.6 µM)
  • α2B (Ki = 2.420 µM)
  • α2C (Ki = 0.860 µM)

The cognitive improvement brought about by MPH could be prevented by α2-adrenoceptor antagonists.74 Guanfacine and clonidine also have a positive effect as α2-adrenoceptor agonists in ADHD.

  • Blockade of the alpha-2-adrenoceptor22
    • In contrast, several sources report an agonistic effect of MPH on the alpha-2-adrenoceptor.7374 See above.

2.1.3. MPH and serotonin

The overall influence of MPH on serotonin levels appears to be negligible.75 D-threo-(R,R)-methylphenidate is a weak agonist of the 5HT-1A serotonin receptor, but not of the 5HT-2A receptor. This can influence the dopamine metabolism in the brain,76 but the extent is small.
MPH is said to bind 2200 times more strongly to the DAT than to the SERT and almost 1300 times more strongly to the NET than to the SERT.41

Whether MPH binds to serotonin receptors is unclear. Various studies have produced contradictory results.41

Controversial:

  • Whether a reuptake inhibition of serotonin occurs at the synapse. There are sources for this77 as well as against it.78
    • The serotonergic effect of MPH is so weak that it is not relevant for the treatment
    • Our impression is that MPH has no significant mood-enhancing effect
2.1.3.1. MPH and tryptophan hydroxylase

Of the two tryptophan hydroxylase isoforms, TPH1 and TPH2, only TPH2 is present in the brain. TPH catalyzes the rate-limiting step in the synthesis of serotonin by converting tryptophan into the serotonin precursor 5-hydroxytryptophan.63
The AATGGAGA (Yin) haplotype of TPH2 appears to be less responsive to MPH than the CGCAAGAC (Yang) haplotype.63

2.1.3.2. Effect of MPH on tryptophan metabolites

In ADHD-HI sufferers (predominantly hyperactive) with comorbid depressive symptoms, one study found significantly higher morning than evening levels of indoleacetic acid compared to ADHD-I sufferers and healthy controls. MPH reduced this by 50 %. MPH also reduced the morning levels of indolepropionic acid and brought the daily profile back to the levels of healthy controls.79

2.1.3.3. Effect of MPH on dorsal raphe nuclei

Several days of MPH administration triggered

  • behavioral sensitization in some rats (which correlated with neuronal arousal) and
  • behavioral tolerance (which was associated with neuronal attenuation) in other rats.
    The neurons of the dorsal raphe nuclei (serotonergic) responded most strongly to acute and chronic MPH administration and differently to the neurons in VTA (dopaminergic) or locus coeruleus (noradrenergic) at all 3 doses used.80
    Dorsal raphe nuclei and serotonin appear to be involved in the acute and chronic effects of MPH and play an independent role in the response to MPH.

2.1.4. Binding affinity of MPH, AMP, ATX to DAT / NET / SERT

The active ingredients methylphenidate (MPH), d-amphetamine (d-AMP), l-amphetamine (l-AMP) and atomoxetine (ATX) bind with different affinities to dopamine transporters (DAT), noradrenaline transporters (NET) and serotonin transporters (SERT). The binding causes an inhibition of the activity of the respective transporters.81

Binding affinity: stronger with smaller number (KD = Ki) DAT NET SERT
MPH 34 - 200 339 > 10,000
d-AMP (Elvanse, Attentin) 34 - 41 23.3 - 38.9 3,830 - 11,000
l-AMP 138 30.1 57,000
ATX 1451 - 1600 2.6 - 5 48 - 77

2.1.5. Effect of MPH, AMP, ATX on dopamine / noradrenaline per brain region

The active substances methylphenidate (MPH), amphetamine (AMP) and atomoxetine (ATX) alter extracellular dopamine (DA) and noradrenaline (NE) to different degrees in different regions of the brain. Table based on Madras,81 modified.

PFC Striatum Nucleus accumbens
MPH DA +
NE (+)
DA +
NE +/- 0
DA +
NE +/- 0
AMP DA +
NE +
DA +
NE +/- 0
DA +
NE +/- 0
ATX DA +
NE +
DA +/- 0
NE +/- 0
DA +/- 0
NE +/- 0

2.1.6. Effect of MPH on MAO-A

MPH influences monoamine oxidase A (MAO-A) by82

  • Stimulation of non-vesicular release
  • Inhibition of MAO-A activity5083

However, the influence appears to be limited and of little relevance.

2.2. Effect of MPH on cholesterol metabolism in the OFC

One study found 12 altered metabolic metabolites in the PFC in SHR rats, considered an ADHD-HI model, compared to WKY rats, considered a model of non-affected individuals. The deviations of 8 of these were equalized by MPH:84

  • 3-Hydroxymethylglutaric acid
  • 3-phosphoglyceric acid
  • Adenosine monophosphate
  • Cholesterol
  • Lanosterine
  • O-Phosphoethanolamine
  • 3-Hydroxymethylglutaric acid
  • Cholesterol

The altered metabolites belong to the metabolic pathways of cholesterol.
In the case of the SHR, the PFC found that

  • Reduced activity of 3-hydroxy-3-methyl-glutaryl-CoA reductase
    • Unchanged by MPH
  • Reduced expression of the sterol regulatory element-binding protein-2
    • Increased by MPH
  • Reduced expression of the ATP-binding cassette transporter A1
    • Increased by MPH

2.3. Effect of MPH on the HPA axis

Stimulants (methylphenidate and amphetamine drugs) are said to increase the activity of the HPA axis.85
MPH increases the cortisol awakening response, which is a sign of reactivity of the HPA axis.86

MPH increased physiological measures of stress (salivary cortisol and blood pressure). MPH modulated the effects of stress on the activation of brain areas associated with goal-directed behavior, including the insula, putamen, amygdala, mPFC, frontal pole, and OFC. However, MPH did not modulate the tendency of stress to cause a reduction in goal-directed behavior.87

2.4. Effect of MPH on the autonomic nervous system (sympathetic / parasympathetic nervous system)

In ADHD, heart rate variability (HRV), which correlates with the health of the autonomic nervous system and in particular reflects the activity of the parasympathetic nervous system, is reduced. Stimulants such as methylphenidate improve (increase) heart rate variability without, however, being able to raise it to the level of non-affected people.8889

The statement made elsewhere,90, that methylphenidate does not change the HVR is not found in the source cited.91

2.5. Effect of MPH on androgens

Stimulants (methylphenidate and amphetamine drugs) reduce the concentration of androgens.
Preclinical data on the role of androgens in the pathogenesis of ADHD suggest that elevated testosterone levels may reduce cerebral blood flow in the PFC by decreasing the amount of alpha estrogen receptors and vascular endothelial growth factor (VEGF). This can interfere with memory processes. There is a correlation between ADHD and the polymorphism of the androgen receptor gene, which leads to its higher expression. Nevertheless, little is known about the question of androgen involvement in ADHD.85

2.6. Effect of MPH on kynurenine

MPH appears to improve the homeostatic ratio of various kynurenines (e.g. increased kynurenic acid vs. decreased quinolinic acid in plasma) in children with ADHD.92

2.7. Effect of MPH on ARAS

Methylphenidate increases the excitation of the reticular activation system (ARAS).93

2.8. Effect of MPH on oxidative/nitrosative status

MPH improved the redox profile with a reduction in the levels of advanced oxidation protein products (AOPP), lipid peroxidation (LPO) and nitrite plus nitrate (NOx) and an increase in the enzymatic activities of glutathione reductase (GRd) and catalase (CAT).86

2.9. Effect of MPH on S100B

One study found that triple therapy (TT) with methylphenidate (MPH), melatonin (aMT) and omega-3 fatty acids (ω-3 PUFAs) increased S100B in ADHD sufferers. The authors see this as an indication that a neuroinflammatory cause of ADHD may damage glial function and thereby alter dopaminergic (DA) neurotransmission.94

2.10. Effect of MPH on brain networks

2.10.1. MPH and connectivity between brain regions

In one study, methylphenidate normalized reduced global connectivity in ADHD 400-700 ms after a stimulus and reduced an increase in network disconnection 100-400 ms after the stimulus. These global changes caused by methylphenidate occurred mainly in the task-relevant frontal and parietal regions and were more significant and lasting than in the non-treated comparison subjects. The results of the study indicate that methylphenidate corrects impaired network flexibility in ADHD.95

Another study reports interhemispheric connectivity changes in ADHD:96

  • Reduced interhemispheric coherence in the delta band in frontal brain regions
  • Increased coherence in the theta band in posterior regions (only with eyes open)
  • Increased coherence in the theta band in central areas

2.10.2. Effect of MPH on Default Mode Network (DMN)

The increased purely intrinsically motivated control of attention in ADHD means that attention and its controllability are just as high as in non-affected individuals when interest is high and only deviates from the attention of non-affected individuals when intrinsic interest is low. This is controlled by the DMN.
Stimulants are able to align the attentional control of ADHD sufferers with that of non-affected persons in the absence of intrinsic interest.97 This explains why stimulants are just as helpful for ADHD-HI and ADHD-C as they are for ADHD-I.

More on the deviant function of the DMN in ADHD and its normalization by stimulants, including further sources, at Normalization of the DMN by stimulants In the article Brain networks and connectivity in ADHD in the chapter Neurological aspects.

2.10.3. Effect of MPH on nucleus accumbens and cognitive control networks

Methylphenidate increased spontaneous neuronal activity in the nucleus accumbens and in cognitive control networks in children with ADHD. This resulted in more stable sustained attention.98

2.11. Effect of MPH on EEG

MPH caused99

  • Significant differences in ADHD sufferers in the frontal-parietal area at 250 ms-400 ms after the stimulus (P3)
  • A decrease in the late 650 ms-800 ms ERP component (LC) at frontal electrodes of ADHD patients compared to controls
  • A significant reduction in reaction time variability in ADHD patients, which correlated with increased P3-ERP response at the frontoparietal electrodes

2.12. Effects on brain regions

Neuroimaging studies show several effects of MPH on different brain regions. These show that MPH acts primarily in the PFC and striatum. MPH

  • Apparently reduces the reduction in gray matter typical of ADHD
    • In the basal ganglia (mainly in children, problem probably decreasing per se in adults)
      • In the right lentiform nucleus100
        • In the right globus pallidus101
        • In the putamen101
      • In the caudate nucleus100
    • In the anterior cingulate cortex (ACC) in adults100
  • MPH mediates its acute and chronic effects on behavior via the dopaminergic system of the caudate nucleus.102
  • In hypermotor and inattentive ADHD sufferers, regular administration of methylphenidate increases the previously unusually low blood flow to the putamen. In ADHD-affected children with average motor activity, regular administration of methylphenidate caused a reduction in blood flow to the putamen. The thalamus was not affected by MPH.103
    MPH increased activation in the bilateral inferior frontal cortex/insula during inhibition of temporal discrimination.104
  • Methylphenidate increases the metabolism in the brain on the left frontal posterior and left parietal superior and decreases it on the left parietal, left parietooccipital and frontal anterior medial.105

MPH appears to reduce dysfunction in the PFC in most affected individuals.106 Another meta-study found that MPH had no effect on working memory (in the dlPFC).104

A study in rats with 0, 0.6, 2.5 and 10.0 mg MPH/kg as single and repeated doses found that MPH acted on the PFC and caudate nucleus. The same dose of MPH induced behavioural sensitization in some animals and tolerance in others, with activity in the PFC and caudate nucleus correlating with the animals’ behavioural responses to MPH. The reaction of the caudate nucleus was more intense than that in the PFC, with both single and repeated administration. In addition, different dose-dependent responses were found between PFC and caudate nucleus: some PFC and caudate nucleus cell units responded to the same MPH dose with excitation and others with attenuation of the neuronal firing rate.107

2.13. MPH for preschool children

Some studies show a positive effect of MPH in preschool children with ADHD.108

2.14. MPH normalizes pain sensation in ADHD

ADHD sufferers often show an increased sensitivity to pain. MPH can eliminate this sensitivity to pain in ADHD sufferers.109

2.15. Serdexmethylphenidate improves sleep in ADHD

One study reports a significant improvement in sleep in children with ADHD between the ages of 6 and 12 with serdexmethylphenidate or dexmethylphenidate.110

2.16. More about MPH

Methylphenidate and amphetamine drugs increase the power of alpha (in rats), while atomoxetine and guanfacine do not.111

MPH acts (among other things) on the dopamine transporters in the brain. As the number of dopamine transporters decreases with increasing age (halving in 50-year-olds compared to 10-year-olds), adults require significantly lower doses.

Details on resumption inhibition

Cranial nerves transmit their information electrically. At a point of contact between a nerve and another nerve (synapse), the signal is passed on to another brain nerve via the synaptic cleft. This transmission of information is usually carried out chemically by neurotransmitters (dopamine, noradrenaline, serotonin and others). The electrical signal causes a release of neurotransmitters (here: dopamine) into the synaptic cleft at the end of the nerve (presynaptic). At the receptor nerve on the other side of the synaptic cleft (postsynaptic), the neurotransmitter (here: dopamine) is taken up by (here: dopamine) receptors and triggers (electrical) signal transmission when a threshold value of activated receptors is reached. The precious neurotransmitter is then returned to the synaptic cleft by the receiving nerve, from where the sending nerve takes up the neurotransmitter again through special reuptake transporters (in the case of dopamine, the dopamine reuptake transporter, DAT) in order to be stored in the vesicles again for the next signal transmission.
In ADHD, the DAT reuptake transporters (primarily located in the striatum) are overactive. If dopamine is released into the synaptic cleft by the transporters of the transmitter nerve, the DAT of the presynaptic transmitter nerve absorb the dopamine again before it can be taken up by the postsynaptic transporters of the receptor nerve. The signal chain is thus disrupted, comparable to the noise of a radio signal (“neural noise”) in relation to dopamine.112 Stimulants such as methylphenidate slow down the activity of the DAT so that the dopamine remains in the synaptic cleft long enough for the signal to be transmitted cleanly. In this way, MPH improves the neural noise in ADHD sufferers to the level of non-affected people.112
The special feature of dopaminergic synapses is that, according to the latest findings (2019), there are no dopamine receptors on the receptor side of the dopaminergic synapse, but rather GABA receptors. Instead, the dopamine receptors are arranged spatially around the synapse and react to dopamine diffusing or otherwise escaping from the synapse.

It is occasionally postulated that very early treatment with stimulants could permanently improve DAT overactivity (i.e. beyond the intake).113

Early medication to cure ADHD?

Early childhood stress exposure leads to long-term damage to the stress regulation systems if there is a corresponding genetic disposition. Such an establishment of stress exposure could possibly be prevented by timely drug treatment. In mice exposed to stress, the serotonin reuptake inhibitor fluoxetine reduced stress-induced increased risk-taking, while the GABA-A receptor agonist diazepam did not.114

Chronic administration of caffeine or MPH before puberty improved object recognition in adult SHR (a rat strain representing a genetic form of ADHD-HI), while the same treatment worsened it in adult Wistar rats115

As the neurotransmitter systems that regulate stress are formed, adjusted and then solidified in the first few years of life (presumably 6 years and earlier), medication that influences this should start much earlier. Whether this will work remains to be seen. What is certain, however, is that child-centered behavioral therapy is of little benefit for small children, while parent-centered therapy brings considerable benefits. This could indicate that the stress systems in young children can still be repaired by external influences.

A very small fMRI study of 16 subjects on the effect of methylphenidate on ADHD-affected and unaffected boys found increased activation of the frontal cortex and reduced activation of the striatum in ADHD-affected boys before taking methylphenidate compared to unaffected boys in go/no-go tasks. Methylphenidate compensated for the differences.116

It remains to be seen what concentrations of methylphenidate reach the synaptic cleft.34
Methylphenidate could accumulate in the central nervous system through active accumulation processes, so that the effective brain concentrations are considerably higher than in plasma.117 For cocaine, the striatal concentration in animals appears to be about 6 times higher than in plasma.118

3. Differences in effect between methylphenidate and amphetamine medication

Methylphenidate may increase left frontal posterior and left parietal superior brain metabolism and decrease left parietal, left parietooccipital and frontal anterior medial metabolism.119

In contrast, D-amphetamine possibly increases metabolism in the right caudate nucleus (part of the striatum) and decreases it in the right Rolandi region and in the right anterior inferior frontal regions.120

The samples (n) on which these findings were examined were very small at 19 and 18. Samples that are too small harbor the considerable risk of misleading results.
Find out more at Studies say - sometimes nothing at all.

4. Effect on symptoms

Methylphenidate improves in children with ADHD:121

  • Response time
  • Response time variability
  • tonic attention
  • phasic attention
  • divided attention
  • Flexibility/shifting attention/task switching
  • selective attention
    • Escapement
    • focused attention
  • Task accuracy in relation to
    • Vigilance
    • divided attention
    • Escapement
    • focused attention
    • Flexibility
    • Obtain integration of sensory information
  • Number of errors of omission and commission in attention tasks

There is evidence for an effect of MPH on neuropathic pain.122

4.1. Particularly good effect of methylphenidate

  • Hyperactivity 93

  • Restlessness93

  • Impulsiveness93

    • Those affected reported in forums that MPH works better against impulsivity than Elvanse.123
    • A study on monkeys (naturally not affected by ADHD) came to the conclusion that low doses of MPH reduced impulsivity, while higher doses had a sedative effect.124
      This follows on from empirical experience that an overdose of MPH can have an apathetic effect.
  • Aggressiveness93125

    • And better than atomoxetine126
    • In a study of 6- to 12-year-old children with aggression and ADHD, systematically titrated stimulants eliminated aggression in 63% of cases.127 In the children in whom stimulants did not sufficiently eliminate aggression, augmentative administration of risperidone (effect size 1.3) or valproic acid (effect size 0.9) improved aggression, with risperidone being associated with weight gain.
  • Socially maladjusted behavior93

  • Behavioral problems, and better than atomoxetine126

  • Somatic complaints, and better than atomoxetine126

  • Motivate through reward128

  • Drive

    • Those affected report quite consistently that MPH improves the drive more than AMP
  • Number of errors of omission and commission in attention tasks129

MPH is effective in adults:130

  • against the core symptoms of ADHD (SMD: 0.49)
  • against the accompanying emotional dysregulation (SMD: 0.34)

MPH is said to work best on the cognitive ADHD symptoms. Motor and social behavior may gradually require slightly higher doses.131

4.2. Good effect of methylphenidate

  • Perception132

  • Concentration93

    • Many adults report that MPH enables greater focus than Elvanse, while Elvanse makes them more relaxed overall and has a more even effect
  • Attention93

    • Distractibility is reduced, attention is increased
    • Task changes are reduced133
  • Motor restlessness93

  • Typeface and graphic expression 134

  • Social perceptiveness and mimic responsiveness

  • Social interaction

    • One study found that basal oxytocin levels in children with ADHD were unchanged compared to those without ADHD. While oxytocin increased in non-affected children after interaction with a parent, oxytocin decreased in untreated ADHD sufferers. Methylphenidate caused the oxytocin increase in ADHD sufferers after parent interaction to correspond to that of non-affected persons.135
  • Rejection sensitivity (offendingness)
    Almost all of the people we interviewed reported an improvement in their rejection sensitivity (which affects almost all of the ADHD sufferers we interviewed) as a result of MPH. A few of those affected reported that their RS became stronger under MPH. One of these sufferers later turned out to be an MPH non-responder who was able to achieve a better effect with an amphetamine medication.

  • Mathematical skills

    • Children with ADHD on MPH showed significantly improved math skills that were indistinguishable from those of non-affected children.136
  • Anxiety125

  • Tension125

  • Borderline aspects125

  • Depressiveness125

  • Emotional instability125

  • Dissatisfaction with life125

  • Negative attitude to life125

  • Psychotic phenomena125

  • Social introversion125

  • Uncertainty125

  • Compulsiveness125

  • Inner emptiness/boredom137

4.3. Low effect of methylphenidate

  • Delay aversion (for adults)138
  • Executive functions (in adults)138

4.4. No effect of methylphendiate

  • Reading the Mind in the Eye (for children). This test measures the theory of mind.139
  • Fine motor skills (handwriting)140

4.5. Different time-dependent effects of stimulants on symptoms?

A publication by a renowned scientist claims different time-response and dose-response curves for the motor and cognitive effects of stimulants.141 While the effect on motor activity lasts 7 to 8 hours, the effect on attention is said to last only 2 to 3 hours. However, the sources cited do not substantiate this claim. Nor do they correspond with empirical experience from practice.

4.6. MPH and smoking cessation

It was reported that sustained-release MPH made a positive contribution to nicotine abstinence / smoking cessation, but only in more severe ADHD cases, whereas in milder ADHD cases a paradoxical worsening occurred, but remitted after discontinuation of the medication.142 This should be seen against the background that nicotine as a stimulant is a self-medication for ADHD, even if smoking uses nicotine as a drug and only nicotine patches or nicotine lozenges act as medication.
Further, in the context of the Inverted-U theory that intermediate neurotransmitter levels mediate optimal brain function, while decreased and elevated neurotransmitter levels cause nearly similar symptoms, the result of this study may indicate overdosing in the subjects with milder ADHD symptoms (indicating lower dopamine and norepinephrine deficiency) and a paradoxical response.

4.7. MPH and creativity

One study found no impairment of creativity by MPH,143 Another study found increased creativity in unmedicated children with ADHD compared to medicated children with ADHD and unaffected children.144

5. Responding (responding / non-responding)

Response means whether there is an effect on the ADHD symptoms. Patients who do not respond sufficiently to a medication are called non-responders.
Non-responding does not mean having no effect, but merely that the effect remains below the level of symptom improvement specified in the respective study.

A meta-study reports a 69% response rate to amphetamine medication and a 59% response rate to methylphenidate. 87% of ADHD sufferers responded to one of the two types of medication.145 A meta-analysis of 32 studies came to the same conclusion (significantly better response rates to amphetamine medication than to MPH).146
For patients for whom MPH does not work, it is therefore advisable to test a medication with amphetamine drugs.
About 50% of patients who do not respond to MPH should respond to atomoxetine, and about 75% of patients who respond to MPH should also respond to atomoxetine.147

In MPH nonresponders, L-amphetamine and atomoxetine were compared in a randomized double-blind study with n = 200 subjects. L-amphetamine was significantly more effective than atomoxetine in 2 of 6 categories and in the overall assessment.148

Positive indications for a response to MPH were:

  • Lower ADHD-RS-IV.es scores149
  • The absence of comorbidities (ODD, depression, alcohol/cannabis use)149
  • Less conspicuous neuropsychological tests149
  • A higher overall IQ149
  • Low commission errors (impulse control errors; reaction to signal that should not have been answered) in the Conners Continuous Performance Test II, CPT-II149
  • Higher hyperactivity-impulsivity and oppositional symptoms before treatment150
    • Predictor for good results with MPH monotherapy, guanfacine monotherapy and MPH/guanfacine combination medication
  • Less anxiety before the treatment150
    • Predictor for good results with MPH monotherapy, guanfacine monotherapy and MPH/guanfacine combination medication
  • High event-related mid-frontal beta power before treatment150
    • EEG activity from cortical sources localized in the regions of the middle frontal bone and middle occipital bone
    • Stronger modulations during encoding and retrieval predictor for good results with MPH monotherapy and guanfacine monotherapy
  • Weak event-related mid-frontal beta power before treatment150
    • EEG activity from cortical sources localized in the regions of the middle frontal bone and middle occipital bone
    • Predictor for good results with MPH/guanfacine combination medication

5.1. Subtypes and non-response probability

Most older sources report that about 90% of sufferers of the ADHD-HI subtype (with hyperactivity) and the mixed type respond positively to methylphenidate and require quite low doses.151152153154155
More recent sources speak of a response rate of up to 75% with MPH,156 which seems more accurate to us.

ADHD-I subtype sufferers are said to be more frequent MPH nonresponders,157 with nonresponder rates of 24%151 being reported. ADHD-I sufferers who respond to MPH also require higher doses.
According to a small study, children with a higher cortisol stress response, which corresponds to the ADHD-I subtype, are more likely to benefit from higher doses of MPH than children with a flattened cortisol stress response (which corresponds to ADHD-HI). However, the stress test was not based on the TSST but on venipuncture, which allows for a less distinct recognizability of the cortisol stress response.158

A particularly strong increase in cortisol awakening (CAR) correlated with reduced MPH responding in children.158

SCT sufferers (which, according to current understanding, is not a subtype of ADHD, but a comorbidity equally common in ADHD-HI and ADHD-I) are particularly frequent MPH nonresponders. In particular, elevated SCT sluggish/sleepy factor values indicate MPH nonresponding. Neither elevated SCT daydreamy symptoms nor ADHD subtype (ADHD-HI or ADHD-I) differed in MPH responding rates in this study.159

According to one study, ADHD sufferers with intellectual deficits are less likely to respond to MPH. A responder rate of 40 to 50 % was reported here.160 In contrast, another study found a good effect of MPH in patients with intellectual deficits.161

5.2. (Non-)responding and EEG subtypes

ADHD sufferers with very low EEG theta values are more likely to be non-responders to stimulants.162
According to this understanding, low theta values correspond to the overactivated beta (EEG) subtype. For the BETA subtype (overactivated type), another source reports reduced MPH responding.48
The beta subtype appears outwardly as the classic ADHD-HI subtype (hyperactive/impulsive). Most people with the ADHD-HI subtype have too low a theta and too high a beta. More on this at ADHD subtypes according to EEG.

However, the (individual) ADHD sufferers of the BETA subtype known to us report an extremely helpful effect of MPH.

A small study found lower EEG stability at rest as a predictor of an MPH response.163
Another study found an attenuated P3 amplitude in responders compared to controls. Unexpectedly, nonresponders showed an atypically flat aperiodic spectral slope compared to controls, while responders did not differ from controls.164

5.3. (Non-)responding and dosing of MPH

Some people suspect cases of underdosing among non-responders, i.e. that the required dosage was not reached and a non-response is only wrongly assumed.165
Our impression is that too low a dosage can cause an apparent non-response. Nevertheless, there are genuine non-responders for whom even greatly increased doses do not produce satisfactory results.
In addition, a different non-responder rate is reported in children and adults.
We suspect that a more precise classification of ADHD subtypes will provide explanations here.

5.4. Indications of good MPH responding

An increase in blood pressure is said to correlate with a particularly good effect of MPH.166

A particularly good improvement in symptoms on methylphenidate was observed in ADHD sufferers with167

  • Increased delta power at F8
  • Increased theta power for Fz, F4, C3, Cz, T5
  • Increased gamma power with T6
  • Reduced beta power at F8 and P3
  • Increased delta/beta power ratio at F8 (in relation to hyperactivity)
  • Increased theta/beta power ratio at F8, F3, Fz, F4, C3, Cz, P3 and T5 (in relation to hyperactivity)

One study found little or no relevance of certain genes that are particularly relevant for neuronal development (“neurodevelopmental network”) to the effect of MPH or atomoxetine in ADHD.168

A meta-analysis of 15 studies and 1382 patients found that carriers of the T allele of the NET gene polymorphism rs28386840 responded significantly more frequently to MPH and showed a significantly greater improvement in hyperactive-impulsive symptoms than carriers of other NET polymorphisms. ADRA2A polymorphisms did not correlate significantly with the response to MPH. However, carriers of the G allele of the MspI polymorphism showed a correlation with a significant improvement in inattention symptoms.169

Elevated iron levels in the putamen and caudate correlated with better MPH responding in ADHD. Elevated iron levels in the putamen correlated - not only in ADHD - with impaired inhibition170

In preschool ADHD, low externalizing or internalizing symptom severity correlated with a high likelihood of response to stimulants. At high externalizing or internalizing symptom levels, the response rate of stimulants approached that of alpha-2 agonists:171

Responder Symptom severity: weak medium strong
Stimulants Externalizing 96.4 % 74.3 % 66.6 %
Alpha-2 agonists Externalizing 40 % 50 % 67 %
Stimulants Internalizing 80.6 % 77.5 % 50 %
Alpha-2 agonists Internalizing 57.7 % 70 % 57.7 %

5.5. Gene variants and MPH effect

5.5.1. ADRA2A gene variants

ADRA2A -1291 polymorphism influences responding and effect of MPH.

  • G/G genotype:
    • 76.9 % responded well to MPH172
    • 25% greater reduction in symptoms after 3 months of MPH173
  • C/G genotype:
    • 46.0 % responded well to MPH172
  • C/G genotype:
    • 41.7 % responded well to MPH172

The genotype of the MspI polymorphism of the ADRA2A gene may influence side effects on OROS MPH:

  • C/C genotype
    • diastolic blood pressure increased by 18.5 % with OROS-MPH174
  • G/G genotype
    • diastolic blood pressure decreased by 0.2 % with OROS-MPH174
  • G/C genotype
    • diastolic blood pressure decreased by 0.2 % with OROS-MPH174

5.5.2. NET gene variants

The genotype of the G1287A polymorphism of the NET gene (noradrenaline transporter, SLC6A2) may influence the response to MPH:

  • G/G genotype:
    • 71.9 % responded well to MPH175
    • 7.15 points improvement on the hyperactivity subscale of the ADHD Rating Scale-IV, but not on the inattention subscale 176
    • no responding difference detected
  • G/A genotype:
    • 46.0 % responded well to MPH175
    • 6.94 points improvement on the hyperactivity subscale of the ADHD Rating Scale-IV, but not on the inattention subscale 176
  • A/A genotype:
    • 57.1 % responded well to MPH175
    • 2.13 points improvement on the hyperactivity subscale of the ADHD Rating Scale-IV, but not on the inattention subscale 176

The genotype of the -3081(A/T) polymorphism of the NET gene (noradrenaline transporter, SLC6A2) may influence the response to MPH:

  • T/T genotype
    • responded comparatively better to MPH177
    • Increase in resting heart rate by 12.5 % due to OROS-MPHi174
  • A/T genotype
    • responded comparatively better to MPH177 -
    • Increase in resting heart rate by 3.5 % due to OROS-MPH174
  • A/A genotype
    • responded comparatively worse to MPH177
    • Increase in resting heart rate by 2.5 % due to OROS-MPH174

A meta-analysis found a correlation between MPH effect and the SLC6A2 gene variants178

  • rs5569 (OR: 1.73)
  • rs28386840 (OR: 2.93)

5.6. CES1 plasma protein level and MPH dosage

Methylphenidate is broken down by the CES1 liver enzyme.
A higher CES1 plasma concentration correlated with a reduced d-methylphenidate plasma level. In one study, the CES1 plasma protein level could explain about 50 % of the variability of the d-methylphenidate plasma level. It is possible that an individualized dosing strategy based on the measurement of CES1 could considerably facilitate the dosing of d-methylphenidate.179

5.7. Response individually dependent on retardation and carrier substance

Those affected report a very different individual response to different MPH preparations.
While the intra-individual (within a person) and inter-individual (compared to other affected persons) differences in tolerability of MPH retard preparations are now generally recognized, it is less well known that tolerability and responding can also be very different intra-individually in relation to different unretarded MPH preparations. We have received reports from a number of patients who reproducibly perceive very clear differences in the effect of various equally strong unretarded MPH preparations.180181


  1. Jaeschke RR, Sujkowska E, Sowa-Kućma M (2021): Methylphenidate for attention-deficit/hyperactivity disorder in adults: a narrative review. Psychopharmacology (Berl). 2021 Oct;238(10):2667-2691. doi: 10.1007/s00213-021-05946-0. PMID: 34436651; PMCID: PMC8455398. REVIEW

  2. Buckner CK, Patil PN, Tye A, Malspeis L (1969): Steric aspects of adrenergic drugs. XII. Some peripheral effects of (+-)-erythro- and (+-)-threo-methylphenidate. J Pharmacol Exp Ther. 1969 Apr;166(2):308-19. PMID: 5813368.

  3. Vadivelu N, Singh-Gill H, Kodumudi G, Kaye AJ, Urman RD, Kaye AD (2014): Practical guide to the management of acute and chronic pain in the presence of drug tolerance for the healthcare practitioner. Ochsner J. 2014 Fall;14(3):426-33. PMID: 25249810; PMCID: PMC4171802.

  4. Arnsten AF, Dudley AG (2005): Methylphenidate improves prefrontal cortical cognitive function through alpha2 adrenoceptor and dopamine D1 receptor actions: Relevance to therapeutic effects in Attention Deficit Hyperactivity Disorder. Behav Brain Funct. 2005 Apr 22;1(1):2. doi: 10.1186/1744-9081-1-2. PMID: 15916700; PMCID: PMC1143775.

  5. Turner PV, Brabb T, Pekow C, Vasbinder MA (2011): Administration of substances to laboratory animals: routes of administration and factors to consider. J Am Assoc Lab Anim Sci. 2011 Sep;50(5):600-13. PMID: 22330705; PMCID: PMC3189662. REVIEW

  6. Gerasimov MR, Franceschi M, Volkow ND, Gifford A, Gatley SJ, Marsteller D, Molina PE, Dewey SL (2000): Comparison between intraperitoneal and oral methylphenidate administration: A microdialysis and locomotor activity study. J Pharmacol Exp Ther. 2000 Oct;295(1):51-7. PMID: 10991960.

  7. Al Shoyaib A, Archie SR, Karamyan VT. Intraperitoneal Route of Drug Administration: Should it Be Used in Experimental Animal Studies? Pharm Res. 2019 Dec 23;37(1):12. doi: 10.1007/s11095-019-2745-x. PMID: 31873819; PMCID: PMC7412579. REVIEW

  8. Thanos PK, Robison LS, Steier J, Hwang YF, Cooper T, Swanson JM, Komatsu DE, Hadjiargyrou M, Volkow ND (2015): A pharmacokinetic model of oral methylphenidate in the rat and effects on behavior. Pharmacol Biochem Behav. 2015 Apr;131:143-53. doi: 10.1016/j.pbb.2015.01.005. PMID: 25641666; PMCID: PMC4461871.

  9. Senior D, Ahmed R, Arnavut E, Carvalho A, Lee WX, Blum K, Komatsu DE, Hadjiargyrou M, Badgaiyan RD, Thanos PK (2023): Behavioral, Neurochemical and Developmental Effects of Chronic Oral Methylphenidate: A Review. J Pers Med. 2023 Mar 23;13(4):574. doi: 10.3390/jpm13040574. PMID: 37108960; PMCID: PMC10144804. REVIEW

  10. Jager A, Kanters D, Geers F, Buitelaar JK, Kozicz T, Glennon JC. Methylphenidate Dose-Dependently Affects Aggression and Improves Fear Extinction and Anxiety in BALB/cJ Mice. Front Psychiatry. 2019 Oct 25;10:768. doi: 10.3389/fpsyt.2019.00768. PMID: 31708820; PMCID: PMC6823535.

  11. Gaytan O, Ghelani D, Martin S, Swann A, Dafny N (1996): Dose response characteristics of methylphenidate on different indices of rats’ locomotor activity at the beginning of the dark cycle. Brain Res. 1996 Jul 15;727(1-2):13-21. doi: 10.1016/0006-8993(96)00296-x. PMID: 8842378.

  12. Senior D, Ahmed R, Arnavut E, Carvalho A, Lee WX, Blum K, Komatsu DE, Hadjiargyrou M, Badgaiyan RD, Thanos PK. Behavioral, Neurochemical and Developmental Effects of Chronic Oral Methylphenidate: A Review. J Pers Med. 2023 Mar 23;13(4):574. doi: 10.3390/jpm13040574. PMID: 37108960; PMCID: PMC10144804. REVIEW; angegebene Fundstellen belegen dies nicht / references given do not prove this

  13. Senior D, Ahmed R, Arnavut E, Carvalho A, Lee WX, Blum K, Komatsu DE, Hadjiargyrou M, Badgaiyan RD, Thanos PK (2023): Behavioral, Neurochemical and Developmental Effects of Chronic Oral Methylphenidate: A Review. J Pers Med. 2023 Mar 23;13(4):574. doi: 10.3390/jpm13040574. PMID: 37108960; PMCID: PMC10144804. REVIEW; angegebene Fundstellen belegen dies nicht / references given do not prove this

  14. Childress, Komolova, Sallee (2019): An update on the pharmacokinetic considerations in the treatment of ADHD with long-acting methylphenidate and amphetamine formulations. Expert Opin Drug Metab Toxicol. 2019 Nov;15(11):937-974. doi: 10.1080/17425255.2019.1675636. PMID: 31581854. REVIEW

  15. Ching, Eslick, Poulton (2019): Evaluation of Methylphenidate Safety and Maximum-Dose Titration Rationale in Attention-Deficit/Hyperactivity Disorder: A Meta-analysis. JAMA Pediatr. 2019 May 28. doi: 10.1001/jamapediatrics.2019.0905.

  16. Krause, Krause (2014): ADHS im Erwachsenenalter: Symptome – Differenzialdiagnose – Therapie, Seite 246

  17. Dodson WW (2005): Pharmacotherapy of adult ADHD. J Clin Psychol. 2005 May;61(5):589-606. doi: 10.1002/jclp.20122. PMID: 15723384. REVIEW

  18. Childress AC, Komolova M, Sallee FR (2019): An update on the pharmacokinetic considerations in the treatment of ADHD with long-acting methylphenidate and amphetamine formulations. Expert Opin Drug Metab Toxicol. 2019 Nov;15(11):937-974. doi: 10.1080/17425255.2019.1675636. PMID: 31581854. REVIEW

  19. Barnhardt EA, Narayanan AR, Coury DL. Evaluating serdexmethylphenidate and dexmethylphenidate capsules as a once-daily treatment option for ADHD. Expert Opin Pharmacother. 2023 May-Aug;24(11):1215-1219. doi: 10.1080/14656566.2023.2218544. PMID: 37226489.

  20. Visentin R, Dalla Man C, Kovatchev B, Cobelli C (2014): The university of Virginia/Padova type 1 diabetes simulator matches the glucose traces of a clinical trial. Diabetes Technol Ther. 2014 Jul;16(7):428-34. doi: 10.1089/dia.2013.0377. PMID: 24571584; PMCID: PMC4074748.

  21. Gutiérrez-Casares JR, Quintero J, Segú-Vergés C, Rodríguez Monterde P, Pozo-Rubio T, Coma M, Montoto C (2023): In silico clinical trial evaluating lisdexamfetamine’s and methylphenidate’s mechanism of action computational models in an attention-deficit/hyperactivity disorder virtual patients’ population. Front Psychiatry. 2023 Jun 2;14:939650. doi: 10.3389/fpsyt.2023.939650. PMID: 37333910; PMCID: PMC10273406.

  22. Steinhausen, Rothenberger, Döpfner (2010): Handbuch ADHS, Seite 84, 85

  23. Berridge CW, Stalnaker TA (2002): Relationship between low-dose amphetamine-induced arousal and extracellular norepinephrine and dopamine levels within prefrontal cortex. Synapse. 2002 Dec 1;46(3):140-9. doi: 10.1002/syn.10131. PMID: 12325041.

  24. Arnsten AF, Dudley AG (2005). Methylphenidate improves prefrontal cortical cognitive function through alpha2 adrenoceptor and dopamine D1 receptor actions: Relevance to therapeutic effects in Attention Deficit Hyperactivity Disorder. Behav Brain Funct. 2005 Apr 22;1(1):2. doi: 10.1186/1744-9081-1-2. PMID: 15916700; PMCID: PMC1143775.

  25. Kuczenski R, Segal DS (2001): Locomotor effects of acute and repeated threshold doses of amphetamine and methylphenidate: relative roles of dopamine and norepinephrine. J Pharmacol Exp Ther. 2001 Mar;296(3):876-83. PMID: 11181919.

  26. Kuczenski R, Segal DS (2002): Exposure of adolescent rats to oral methylphenidate: preferential effects on extracellular norepinephrine and absence of sensitization and cross-sensitization to methamphetamine. J Neurosci. 2002 Aug 15;22(16):7264-71. doi: 10.1523/JNEUROSCI.22-16-07264.2002. PMID: 12177221; PMCID: PMC6757883.

  27. Bymaster FP, Katner JS, Nelson DL, et al. Atomoxetine increases extracellular levels of norepinephrine and dopamine in prefrontal cortex of rat: a potential mechanism for efficacy in attention deficit/hyperactivity disorder. Neuropsychopharmacology. 2002;27:699–711.

  28. Di Miceli, Derf, Gronier (2022): Consequences of Acute or Chronic Methylphenidate Exposure Using Ex Vivo Neurochemistry and In Vivo Electrophysiology in the Prefrontal Cortex and Striatum of Rats. Int J Mol Sci. 2022 Aug 2;23(15):8588. doi: 10.3390/ijms23158588. PMID: 35955717; PMCID: PMC9369023.

  29. Koda K, Ago Y, Cong Y, Kita Y, Takuma K, Matsuda T (2010): Effects of acute and chronic administration of atomoxetine and methylphenidate on extracellular levels of noradrenaline, dopamine and serotonin in the prefrontal cortex and striatum of mice. J Neurochem. 2010 Jul;114(1):259-70. doi: 10.1111/j.1471-4159.2010.06750.x. PMID: 20403082.

  30. Berridge CW, Devilbiss DM, Andrzejewski ME, Arnsten AF, Kelley AE, Schmeichel B, Hamilton C, Spencer RC (2006):Methylphenidate preferentially increases catecholamine neurotransmission within the prefrontal cortex at low doses that enhance cognitive function. Biol Psychiatry. 2006 Nov 15;60(10):1111-20. doi: 10.1016/j.biopsych.2006.04.022. PMID: 16806100.

  31. Kodama T, Kojima T, Honda Y, Hosokawa T, Tsutsui KI, Watanabe M (2017): Oral Administration of Methylphenidate (Ritalin) Affects Dopamine Release Differentially Between the Prefrontal Cortex and Striatum: A Microdialysis Study in the Monkey. J Neurosci. 2017 Mar 1;37(9):2387-2394. doi: 10.1523/JNEUROSCI.2155-16.2017. PMID: 28154152; PMCID: PMC6596846.

  32. Gatzke-Kopp, Beauchaine (2007): Central nervous system substrates of impulsivity: Implications for the development of attention-deficit/hyperactivity disorder and conduct disorder. In: Coch, Dawson, Fischer ( Eds): Human behavior, learning, and the developing brain: Atypical development. New York: Guilford Press; 2007. pp. 239–263; 245

  33. Ruskin DN, Bergstrom DA, Shenker A, Freeman LE, Baek D, Walters JR (2001): Drugs used in the treatment of attention-deficit/hyperactivity disorder affect postsynaptic firing rate and oscillation without preferential dopamine autoreceptor action. Biol Psychiatry. 2001 Feb 15;49(4):340-50. doi: 10.1016/s0006-3223(00)00987-2. PMID: 11239905.

  34. Federici M, Latagliata EC, Ledonne A, Rizzo FR, Feligioni M, Sulzer D, Dunn M, Sames D, Gu H, Nisticò R, Puglisi-Allegra S, Mercuri NB (2014): Paradoxical abatement of striatal dopaminergic transmission by cocaine and methylphenidate. J Biol Chem. 2014 Jan 3;289(1):264-74. doi: 10.1074/jbc.M113.495499. PMID: 24280216; PMCID: PMC3879549.

  35. Fuller, Burrell, Yee, Liyanagama, Lipski, Wickens, Hyland (2019): Role of homeostatic feedback mechanisms in modulating methylphenidate actions on phasic dopamine signaling in the striatum of awake behaving rats. Prog Neurobiol. 2019 Nov;182:101681. doi: 10.1016/j.pneurobio.2019.101681. PMID: 31412279.

  36. Evers EA, Stiers P, Ramaekers JG (2017): High reward expectancy during methylphenidate depresses the dopaminergic response to gain and loss. Soc Cogn Affect Neurosci. 2017 Feb 1;12(2):311-318. doi: 10.1093/scan/nsw124. PMID: 27677943; PMCID: PMC5390715.

  37. Zhang, Zhang, Liang, Siapas, Zhou, Dani (2009): Dopamine signaling differences in the nucleus accumbens and dorsal striatum exploited by nicotine. J Neurosci. 2009 Apr 1;29(13):4035-43. doi: 10.1523/JNEUROSCI.0261-09.2009. PMID: 19339599; PMCID: PMC2743099.

  38. Stahl (2013): Stahl’s Essential Psychopharmacology, 4. Auflage, Chapter 12: Attention deficit hyperactivity disorder and its treatment, Seite 490

  39. Seeman P, Madras BK (1998): Anti-hyperactivity medication: methylphenidate and amphetamine. Mol Psychiatry. 1998 Sep;3(5):386-96. doi: 10.1038/sj.mp.4000421. PMID: 9774771. REVIEW

  40. Hauger, Angel, Janowsky, Berger, Hulihan-Giblin (1990): Brain Recognition Sites for Methylphenidate and the Amphetamines; Their Relationship to the Dopamine Transport Complex, Glucoreceptors, and Serotonergic Neurotransmission in the Central Nervous System; in: Application of Basic Neuroscience to Child Psychiatry, Seiten 77-100

  41. Jaeschke, Sujkowska, Sowa-Kućma (2021): Methylphenidate for attention-deficit/hyperactivity disorder in adults: a narrative review. Psychopharmacology (Berl). 2021 Oct;238(10):2667-2691. doi: 10.1007/s00213-021-05946-0. PMID: 34436651; PMCID: PMC8455398. REVIEW

  42. Gatley, Pan, Chen, Chaturvedi, Ding (1996): Affinities of methylphenidate derivatives for dopamine, norepinephrine and serotonin transporters. Life Sci. 1996;58(12):231-9. doi: 10.1016/0024-3205(96)00052-5. PMID: 8786705.

  43. Krause, la Fougere, Krause, Ackenheil, Dresel (2005): Influence of striatal dopamine transporter availability on the response to methylphenidate in adult patients with ADHD; European Archives of Psychiatry and Clinical Neuroscience; December 2005, Volume 255, Issue 6, pp 428–431, Achtung, geringe Probandenzahl von n = 18

  44. Wang, Volkow, Wigal, Kollins, Newcorn, Telang, Logan, Jayne, Wong, Han, Fowler, Zhu, Swanson (2013): Long-Term Stimulant Treatment Affects Brain Dopamine Transporter Level in Patients with Attention Deficit Hyperactive Disorder. PLoS ONE 8(5): e63023. https://doi.org/10.1371/journal.pone.0063023

  45. Takamatsu, Hagino, Sato, Takahashi, Nagasawa, Kubo, Mizuguchi, Uhl, Sora, Ikeda (2015):Improvement of Learning and Increase in Dopamine Level in the Frontal Cortex by Methylphenidate in Mice Lacking Dopamine Transporter; 1 Curr Mol Med. 2015 Mar; 15(3): 245–252. doi: 10.2174/1566524015666150330144018, PMCID: PMC5384353

  46. Federici, Latagliata, Ledonne, Rizzo, Feligioni, Sulzer, Dunn, Sames, Gu, Nisticò, Puglisi-Allegra, Mercuri (2014): Paradoxical abatement of striatal dopaminergic transmission by cocaine and methylphenidate. J Biol Chem. 2014 Jan 3;289(1):264-74. doi: 10.1074/jbc.M113.495499. PMID: 24280216; PMCID: PMC3879549.

  47. Diamond: Attention-deficit disorder (attention-deficit/hyperactivity disorder without hyperactivity): A neurobiologically and behaviorally distinct disorder from attention-deficit (with hyperactivity), Development and Psychopathology 17 (2005), 807–825, Seite 811

  48. Müller, Candrian, Kropotov (2011), ADHS – Neurodiagnostik in der Praxis, Springer

  49. Porrino, Lucignani (1987): Different patterns of local brain energy metabolism associated with high and low doses of methylphenidate. Relevance to its action in hyperactive children.Biol Psychiatry. 1987 Feb;22(2):126-38.

  50. Solanto (1998): Neuropsychopharmacological mechanisms of stimulant drug action in attention-deficit hyperactivity disorder: a review and integration. Behav Brain Res. 1998 Jul;94(1):127-52. doi: 10.1016/s0166-4328(97)00175-7. PMID: 9708845. REVIEW

  51. https://www.kinderaerzte-im-netz.de/media/53ec94e833af614b730097d1/source/20080530092715_adhs2.pdf

  52. Markowitz, Patrick (2001): Pharmacokinetic and pharmacodynamic drug interactions in the treatment of attention-deficit hyperactivity disorder. Clin Pharmacokinet. 2001;40(10):753-72

  53. Markowitz, Patrick (2001): Pharmacokinetic and pharmacodynamic drug interactions in the treatment of attention-deficit hyperactivity disorder. Clin Pharmacokinet. 2001;40(10):753-72 unter Verweis auf Volkow, Ding, Fowler, Wang, Logan, Gatley, Dewey, Ashby, Liebermann, Hitzemann, et al. (1995): Is methylphenidate like cocaine? Studies on their pharmacokinetics and distribution in the human brain. Arch Gen Psychiatry. 1995 Jun;52(6):456-63 und Hitri, Hurd, Wyatt, Deutsch (1994): Molecular, functional and biochemical characteristics of the dopamine transporter: regional differences and clinical relevance. Clin Neuropharmacol 1994; 17: 1-22

  54. Steinhausen, Rothenberger, Döpfner (2020): Handbuch ADHS 2. Aufl., Seite 92

  55. Volkow, Wang, Fowler, Logan, Angrist, Hitzemann, Lieberman, Pappas (1997): Effects of methylphenidate on regional brain glucose metabolism in humans: relationship to dopamine D2 receptors.Am J Psychiatry. 1997 Jan;154(1):50-5, sehr niedrige Probandenzahl von n = 15

  56. Steinhausen, Rothenberger, Döpfner (2010): Handbuch ADHS, Seite 78

  57. Roessner, Sagvolden, Dasbanerjee, Middleton, Faraone, Walaas, Becker, Rothenberger, Bock (2010): Neuroscience. 2010 Jun 2;167(4):1183-91. doi: 10.1016/j.neuroscience.2010.02.073.

  58. Quansah, Ruiz-Rodado, Grootveld, Zetterström (2018): Methylphenidate alters monoaminergic and metabolic pathways in the cerebellum of adolescent rats. Eur Neuropsychopharmacol. 2018 Feb 22. pii: S0924-977X(18)30043-9. doi: 10.1016/j.euroneuro.2018.02.002.

  59. Quansah, Zetterström (2019): Chronic methylphenidate preferentially alters catecholamine protein targets in the parietal cortex and ventral striatum. Neurochem Int. 2019 Jan 17;124:193-199. doi: 10.1016/j.neuint.2019.01.016.

  60. Kim, Heo, Kim, Ko, Lee, Kim, Kim, Kim, Ji, Kim, Shin, Choi, Kim (2011): Treadmill exercise and methylphenidate ameliorate symptoms of attention deficit/hyperactivity disorder through enhancing dopamine synthesis and brain-derived neurotrophic factor expression in spontaneous hypertensive rats. Neurosci Lett. 2011 Oct 17;504(1):35-9. doi: 10.1016/j.neulet.2011.08.052. PMID: 21907264.

  61. Panos, O’Callaghan, Miller, Ferguson (2014): Effects of developmental methylphenidate (MPH) treatment on monoamine neurochemistry of male and female rats. Neurotoxicol Teratol. 2014 Sep-Oct;45:70-4. doi: 10.1016/j.ntt.2014.08.001. PMID: 25132048.

  62. Bartl, Palazzesi, Parrinello, Hommers, Riederer, Walitza, Grünblatt (2017): The impact of methylphenidate and its enantiomers on dopamine synthesis and metabolism in vitro. Prog Neuropsychopharmacol Biol Psychiatry. 2017 Oct 3;79(Pt B):281-288. doi: 10.1016/j.pnpbp.2017.07.002. PMID: 28690202.

  63. Stevens, Sangkuhl, Brown, Altman, Klein (2019): PharmGKB summary: methylphenidate pathway, pharmacokinetics/pharmacodynamics. Pharmacogenet Genomics. 2019 Aug;29(6):136-154. doi: 10.1097/FPC.0000000000000376. PMID: 30950912; PMCID: PMC6581573.

  64. Uchimura H, Kanai K, Arai M, Inoue M, Hishimoto A, Masukawa D, Goshima Y (2023): Involvement of the L-DOPA receptor GPR143 in acute and chronic actions of methylphenidate. J Pharmacol Sci. 2023 Jul;152(3):178-181. doi: 10.1016/j.jphs.2023.04.006. PMID: 37257945.

  65. Berridge, Devilbiss, Andrzejewski, Arnsten, Kelley, Schmeichel, Hamilton, Spencer (2006): Methylphenidate Preferentially Increases Catecholamine Neurotransmission within the Prefrontal Cortex at Low Doses that Enhance Cognitive Function; Biological Psychiatry; Volume 60, Issue 10, 15 November 2006, Pages 1111-1120

  66. Koda, Ago, Cong, Kita, Takuma, Matsuda (2010): Effects of acute and chronic administration of atomoxetine and methylphenidate on extracellular levels of noradrenaline, dopamine and serotonin in the prefrontal cortex and striatum of mice. J Neurochem. 2010 Jul;114(1):259-70. doi: 10.1111/j.1471-4159.2010.06750.x. PMID: 20403082.

  67. Ruocco, Carnevale, Treno, Sadile, Melisi, Arra, Ibba, Schirru, Carboni (2010): Prepuberal subchronic methylphenidate and atomoxetine induce different long-term effects on adult behaviour and forebrain dopamine, norepinephrine and serotonin in Naples high-excitability rats. Behav Brain Res. 2010 Jun 26;210(1):99-106. doi: 10.1016/j.bbr.2010.02.020.

  68. Mulvihill (2019): Presynaptic regulation of dopamine release: Role of the DAT and VMAT2 transporters. Neurochem Int. 2019 Jan;122:94-105. doi: 10.1016/j.neuint.2018.11.004. PMID: 30465801. REVIEW

  69. Lam, Matthies, Graf, Colla, Jacob, Sobanski, Alm, Rösler, Retz, Retz-Junginger, Kis, Abdel-Hamid, Müller, Lücke, Huss, Jans, Berger, Tebartz van Elst, Philipsen (2019): Comparison of Methylphenidate and Psychotherapy in Adult ADHD Study (COMPAS) Consortium. Long-term Effects of Multimodal Treatment on Adult Attention-Deficit/Hyperactivity Disorder Symptoms: Follow-up Analysis of the COMPAS Trial. JAMA Netw Open. 2019 May 3;2(5):e194980. doi: 10.1001/jamanetworkopen.2019.4980. PMID: 31150084; PMCID: PMC6547099.

  70. Alam, Wasi, Naeem, Kashif, Siddiqui, Bashir, Naz, Ikram (2019): Methylphenidate increases the urinary excretion of vanillylmandelic acid in rats that is attenuated by buspirone co-administration. Pak J Pharm Sci. 2019 Mar;32(2):895-898.

  71. Levi-Sachar, Gvirts, Goldwin, Bloch, Shamay-Tsoory, Zagoory-Sharon, Feldman, Maoz (2019): The effect of methylphenidate on social cognition and oxytocin in children with attention deficit hyperactivity disorder. Neuropsychopharmacology. 2019 Sep 12. doi: 10.1038/s41386-019-0522-5. n = 90

  72. Kuczenski R, Segal DS (1997): Effects of methylphenidate on extracellular dopamine, serotonin, and norepinephrine: comparison with amphetamine. J Neurochem. 1997 May;68(5):2032-7. doi: 10.1046/j.1471-4159.1997.68052032.x. Erratum in: J Neurochem 1997 Sep;69(3):1332. PMID: 9109529.

  73. Faraone (2018): The pharmacology of amphetamine and methylphenidate: Relevance to the neurobiology of attention-deficit/hyperactivity disorder and other psychiatric comorbidities. Neurosci Biobehav Rev. 2018 Apr;87:255-270. doi: 10.1016/j.neubiorev.2018.02.001. PMID: 29428394.

  74. Gamo, Wang, Arnsten (2010): Methylphenidate and atomoxetine enhance prefrontal function through α2-adrenergic and dopamine D1 receptors. J Am Acad Child Adolesc Psychiatry. 2010 Oct;49(10):1011-23. doi: 10.1016/j.jaac.2010.06.015. PMID: 20855046; PMCID: PMC2999884.

  75. Dinis-Oliveira (2017): Metabolomics of Methylphenidate and Ethylphenidate: Implications in Pharmacological and Toxicological Effects. Eur J Drug Metab Pharmacokinet. 2017 Feb;42(1):11-16. doi: 10.1007/s13318-016-0362-1. PMID: 27438788. REVIEW

  76. Markowitz, DeVane, Ramamoorthy, Zhu (2009): The psychostimulant d-threo-(R,R)-methylphenidate binds as an agonist to the 5HT(1A) receptor. Pharmazie. 2009 Feb;64(2):123-5.

  77. Garattini, Mennini (1988): Critical Notes on The Specificity of Drugs in The Study of Metabolism and Functions of Brain Monoamines; International Review of Neurobiology; Volume 29, 1988, Pages 259–280, http://dx.doi.org/10.1016/S0074-7742(08)60089-6, zitiert nach Trott, Wirth (2000): die Pharmakotherapie der hyperkinetischen Störungen; in: Steinhausen (Herausgeber): Hyperkinetischen Störungen bei Kindern, Jugendlichen und Erwachsenen, 2. Aufl., Seite 210 REVIEW

  78. Leonard, McCartan, White, King (2004): Methylphenidate: a review of its neuropharmacological, neuropsychological and adverse clinical effects. Hum. Psychopharmacol. Clin. Exp., 19: 151–180. REVIEW

  79. Fernández-Lópe, Molina-Carballo, Cubero-Millán, Checa-Ros, Machado-Casas, Blanca-Jover, Jerez-Calero, Madrid-Fernández, Uberos, Muñoz-Hoyos (2020): Indole Tryptophan Metabolism and Cytokine S100B in Children with Attention-Deficit/Hyperactivity Disorder: Daily Fluctuations, Responses to Methylphenidate, and Interrelationship with Depressive Symptomatology. J Child Adolesc Psychopharmacol. 2020 Feb 12. doi: 10.1089/cap.2019.0072. PMID: 32048862.

  80. Yuan A, Claussen C, Jones Z, Tang B, Dafny N (2023): Methylphenidate induces a different response in the dorsal raphe as compared to ventral tegmental area and locus coeruleus: behavioral and concomitant neuronal recordings in adult rats. J Neural Transm (Vienna). 2023 Jul 1. doi: 10.1007/s00702-023-02665-y. PMID: 37391573.

  81. Madras, Miller, Fischman (2005): The dopamine transporter and attention-deficit/hyperactivity disorder. Biol Psychiatry. 2005 Jun 1;57(11):1397-409. doi: 10.1016/j.biopsych.2004.10.011. PMID: 15950014.

  82. Masellis, Basile, Kennedy (2006): Neuropsychopharmacogenetics: ‘stimulating’ rationale therapy in attention-deficit/hyperactivity disorder (ADHD): pharmacogenetics of psychostimulants in ADHD. In: Gorwood, Hamon (editors): Psychopharmacogenetics. Boston: Springer US; 2006. p. 231 - 248.

  83. Mefford, Potter (1989): A neuroanatomical and biochemical basis for attention deficit disorder with hyperactivity in children: a defect in tonic adrenaline mediated inhibition of locus coeruleus stimulation; Med Hypotheses. 1989 May;29(1):33-42., zitiert nach Trott, Wirth (2000): die Pharmakotherapie der hyperkinetischen Störungen; in: Steinhausen (Herausgeber) hyperkinetischen Störungen bei Kindern, Jugendlichen und Erwachsenen, 2. Aufl., Seite 210

  84. Chen, Yuan, Sun, Song, Lu, Ni, Han (2019): Metabolomics study of the prefrontal cortex in a rat model of attention deficit hyperactivity disorder reveals the association between cholesterol metabolism disorder and hyperactive behavior. Biochem Biophys Res Commun. 2019 Dec 18. pii: S0006-291X(19)32336-8. doi: 10.1016/j.bbrc.2019.12.016.

  85. Budziszewska, Basta-Kaim, Kubera, Lasoń (2010); [Immunological and endocrinological pattern in ADHD etiopathogenesis]. Przeglad Lekarski [01 Jan 2010, 67(11):1200-1204], PMID:21442976

  86. Garre-Morata L, de Haro T, Villén RG, Fernández-López ML, Escames G, Molina-Carballo A, Acuña-Castroviejo D (2024): Changes in Cortisol and in Oxidative/Nitrosative Stress Indicators after ADHD Treatment. Antioxidants (Basel). 2024 Jan 12;13(1):92. doi: 10.3390/antiox13010092. PMID: 38247516; PMCID: PMC10812591.

  87. van Ruitenbeek, Quaedflieg, Hernaus, Hartogsveld, Smeets (2021): Dopaminergic and noradrenergic modulation of stress-induced alterations in brain activation associated with goal-directed behaviour. J Psychopharmacol. 2021 Sep 14:2698811211044679. doi: 10.1177/02698811211044679. PMID: 34519561.

  88. Negrao, Crafford, Viljoen (2009): The effect of sympathomimetic medication on cardiovascular functioning of children with attention-deficit/hyperactivity disorder. Cardiovasc J Afr. 2009;20(5):296-9

  89. Hammerness, Wilens, Mick, Spencer, Doyle, McCreary, Becker, Biederman (2009): Cardiovascular effects of longer-term, high-dose OROS methylphenidate in adolescents with attention deficit hyperactivity disorder. J Pediatr. 2009;155(1):84-9, 89.e1

  90. von http://www.hrv24.de/HRV-Medikamente.htm

  91. Vetter, Elia, Erickson, Berger, Blum, Uzark, Webb (2008): Cardiovascular Monitoring of Children and Adolescents With Heart Disease Receiving Medications for Attention Deficit/Hyperactivity Disorder. A Scientific Statement From the American Heart Association Council on Cardiovascular Disease in the Young Congenital Cardiac Defects Committee and the Council on Cardiovascular Nursing. Circulation. 2008;117:2407-23.

  92. Molina-Carballo, Cubero-Millán, Fernández-López, Checa-Ros, Machado-Casas, Jerez-Calero, Blanca-Jover, Cantarero-Malagón, Uberos, Muñoz-Hoyos (2021): Methylphenidate ameliorates the homeostatic balance between levels of kynurenines in ADHD children. Psychiatry Res. 2021 Jun 17;303:114060. doi: 10.1016/j.psychres.2021.114060. PMID: 34175711. n = 179

  93. Trott, Wirth (2000): die Pharmakotherapie der hyperkinetischen Störungen; in: Steinhausen (Herausgeber) hyperkinetischen Störungen bei Kindern, Jugendlichen und Erwachsenen, 2. Aufl., Seite 211

  94. Ouadih-Moran M, Muñoz-Hoyos A, D’Marco L, Molina-Carballo A, Seiquer I, Checa-Ros A (2023): Is S100B Involved in Attention-Deficit/Hyperactivity Disorder (ADHD)? Comparisons with Controls and Changes Following a Triple Therapy Containing Methylphenidate, Melatonin and ω-3 PUFAs. Nutrients. 2023 Jan 31;15(3):712. doi: 10.3390/nu15030712. PMID: 36771418.

  95. Rubinson, Horowitz, Naim-Feil, Gothelf, Moses, Levit-Binnun (2019): EEG functional networks reveal global effects of Methylphenidate in youth with Attention Deficit/Hyperactivity Disorder. Brain Connect. 2019 Mar 28. doi: 10.1089/brain.2018.0630.

  96. Markovska-Simoska, Pop-Jordanova, Pop-Jordanov (2019): Inter- and Intra-Hemispheric EEG Coherence Study in Adults with Neuropsychiatric Disorders. Pril (Makedon Akad Nauk Umet Odd Med Nauki). 2018 Dec 1;39(2-3):5-19. doi: 10.2478/prilozi-2018-0037.

  97. Liddle, Hollis, Batty, Groom, Totman, Liotti, Scerif, Liddle (2011): Task-related default mode network modulation and inhibitory control in ADHD: effects of motivation and methylphenidate. J Child Psychol Psychiatry. 2011 Jul;52(7):761-71. doi: 10.1111/j.1469-7610.2010.02333.x.

  98. Mizuno Y, Cai W, Supekar K, Makita K, Takiguchi S, Silk TJ, Tomoda A, Menon V (2022): Methylphenidate Enhances Spontaneous Fluctuations in Reward and Cognitive Control Networks in Children With Attention-Deficit/Hyperactivity Disorder. Biol Psychiatry Cogn Neurosci Neuroimaging. 2022 Oct 23:S2451-9022(22)00247-6. doi: 10.1016/j.bpsc.2022.10.001. PMID: 36717325.

  99. Rubinson, Horowitz, Naim-Feil, Gothelf, Levit-Binnun, Moses (2019): Effects of methylphenidate on the ERP amplitude in youth with ADHD: A double-blind placebo-controlled cross-over EEG study. PLoS One. 2019 May 31;14(5):e0217383. doi: 10.1371/journal.pone.0217383. eCollection 2019.

  100. Nakao, Radua, Rubia, Mataix-Cols (2011): Gray matter volume abnormalities in ADHD: voxel-based meta-analysis exploring the effects of age and stimulant medication. Am J Psychiatry. 2011 Nov;168(11):1154-63. doi: 10.1176/appi.ajp.2011.11020281. 14 Studien mit n = 722

  101. Frodl, Skokauskas (2012): Meta-analysis of structural MRI studies in children and adults with attention deficit hyperactivity disorder indicates treatment effects. Acta Psychiatr Scand. 2012 Feb;125(2):114-26. doi: 10.1111/j.1600-0447.2011.01786.x.

  102. King, Floren, Kharas, Thomas, Dafny (2019): Glutaminergic signaling in the caudate nucleus is required for behavioral sensitization to methylphenidate. Pharmacol Biochem Behav. 2019 Sep;184:172737. doi: 10.1016/j.pbb.2019.172737.

  103. Teicher, Anderson, Polcari, Glod, Maas, Renshaw (2000): Functional deficits in basal ganglia of children with attention-deficit/hyperactivity disorder shown with functional magnetic resonance imaging relaxometry.Nat Med. 2000 Apr;6(4):470-3.

  104. Rubia, Alegria, Cubillo, Smith, Brammer, Radua (2014): Effects of stimulants on brain function in attention-deficit/hyperactivity disorder: a systematic review and meta-analysis. Biol Psychiatry. 2014 Oct 15;76(8):616-28. doi: 10.1016/j.biopsych.2013.10.016. 14 Studien mit n = 222 REVIEW

  105. Matochik, Liebenauer, King, Szymanski, Cohen, Zametkin (1994): Cerebral glucose metabolism in adults with attention deficit hyperactivity disorder after chronic stimulant treatment. Am J Psychiatry 151: 658–664; Achtung, geringes N von 37; Vorsicht, geringe Stichprobe mit n = 19; zitiert nach Edel, Vollmoeller (2006): Aufmerksamkeitsdefizit-/Hyperaktivitätsstörung bei Erwachsenen, Springer, Seite 10

  106. Spencer, Brown, Seidman, Valera, Makris, Lomedico, Faraone, Biederman (2013) Effect of psychostimulants on brain structure and function in ADHD: a qualitative literature review of magnetic resonance imaging-based neuroimaging studies. J Clin Psychiatry. 2013 Sep;74(9):902-17. doi: 10.4088/JCP.12r08287. REVIEW

  107. Venkataraman, Claussen, Kharas, Dafny (2020): The prefrontal cortex and the caudate nucleus respond conjointly to methylphenidate (Ritalin). Concomitant behavioral and neuronal recording study. Brain Res Bull. 2020 Jan 24;157:77-89. doi: 10.1016/j.brainresbull.2019.10.009. PMID: 31987926.

  108. Wigal, Chappell, Palumbo, Lubaczewski, Ramaker, Abbas (2020): Diagnosis and Treatment Options for Preschoolers with Attention-Deficit/Hyperactivity Disorder. J Child Adolesc Psychopharmacol. 2020 Jan 20;10.1089/cap.2019.0116. doi: 10.1089/cap.2019.0116. PMID: 31967914.

  109. Treister R, Eisenberg E, Demeter N, Pud D (2015): Alterations in pain response are partially reversed by methylphenidate (Ritalin) in adults with attention deficit hyperactivity disorder (ADHD). Pain Pract. 2015 Jan;15(1):4-11. doi: 10.1111/papr.12129. PMID: 24134430.

  110. Mattingly GW, Childress AC, Cutler AJ, Estrada J, Corliss M (2023): Serdexmethylphenidate/dexmethylphenidate effects on sleep in children with attention-deficit/hyperactivity disorder. Front Psychiatry. 2023 Jun 23;14:1193455. doi: 10.3389/fpsyt.2023.1193455. PMID: 37426086; PMCID: PMC10327472.

  111. Takahashi, Ohmiya, Honda, Ni (2018): The KCNH3 inhibitor ASP2905 shows potential in the treatment of attention deficit/hyperactivity disorder. PLoS One. 2018 Nov 21;13(11):e0207750. doi: 10.1371/journal.pone.0207750. eCollection 2018.

  112. Pertermann, Bluschke, Roessner, Beste (2019): The Modulation of Neural Noise Underlies the Effectiveness of Methylphenidate Treatment in Attention-Deficit/Hyperactivity Disorder. Biol Psychiatry Cogn Neurosci Neuroimaging. 2019 Apr 3. pii: S2451-9022(19)30080-1. doi: 10.1016/j.bpsc.2019.03.011.

  113. Edel, Vollmoeller (2006): Aufmerksamkeitsdefizit-/Hyperaktivitätsstörung bei Erwachsenen, Springer, Seite 13

  114. Wang, Li, Sawmiller, Fan, Ma, Tan, Ren, Li (2013): Chronic mild stress-induced changes of risk assessment behaviors in mice are prevented by chronic treatment with fluoxetine but not diazepam.Pharmacol Biochem Behav. 2014 Jan;116:116-28. doi: 10.1016/j.pbb.2013.11.028.

  115. Pires, Pamplona, Pandolfo, Prediger, Takahashi (2010): Chronic caffeine treatment during prepubertal period confers long-term cognitive benefits in adult spontaneously hypertensive rats (SHR), an animal model of attention deficit hyperactivity disorder (ADHD). Behav Brain Res. 2010 Dec 20;215(1):39-44. doi: 10.1016/j.bbr.2010.06.022. PMID: 20600342.

  116. Vaidya, Austin, Kirkorian, Ridlehuber, Desmond, Glover, Gabrieli (1998): Selective effects of methylphenidate in attention deficit hyperactivity disorder: a functional magnetic resonance study. Proc Natl Acad Sci U S A. 1998 Nov 24;95(24):14494-9.

  117. Balcioglu A, Ren JQ, McCarthy D, Spencer TJ, Biederman J, Bhide PG (2009): Plasma and brain concentrations of oral therapeutic doses of methylphenidate and their impact on brain monoamine content in mice. Neuropharmacology. 2009 Dec;57(7-8):687-93. doi: 10.1016/j.neuropharm.2009.07.025. PMID: 19631228; PMCID: PMC2783341.

  118. Bystrowska B, Adamczyk P, Moniczewski A, Zaniewska M, Fuxe K, Filip M (2012): LC/MS/MS evaluation of cocaine and its metabolites in different brain areas, peripheral organs and plasma in cocaine self-administering rats. Pharmacol Rep. 2012;64(6):1337-49. doi: 10.1016/s1734-1140(12)70931-3. PMID: 23406744.

  119. Matochik, Liebenauer, King, Szymanski, Cohen, Zametkin (1994): Cerebral glucose metabolism in adults with attention defi cit hyperactivity disorder after chronic stimulant treatment. Am J Psychiatry 151: 658–664; n = 18; zitiert nach Edel, Vollmoeller (2006): Aufmerksamkeitsdefizit-/Hyperaktivitätsstörung bei Erwachsenen, Springer, Seite 10

  120. Matochik, Liebenauer, King, Szymanski, Cohen, Zametkin (1994): Cerebral glucose metabolism in adults with attention defi cit hyperactivity disorder after chronic stimulant treatment. Am J Psychiatry 151: 658–664; geringe Stichprobe mit n = 19; zitiert nach Edel, Vollmoeller (2006): Aufmerksamkeitsdefizit-/Hyperaktivitätsstörung bei Erwachsenen, Springer, Seite 10

  121. Pitzianti MB, Spiridigliozzi S, Bartolucci E, Esposito S, Pasini A (2020): New Insights on the Effects of Methylphenidate in Attention Deficit Hyperactivity Disorder. Front Psychiatry. 2020 Sep 30;11:531092. doi: 10.3389/fpsyt.2020.531092. PMID: 33132928; PMCID: PMC7561436. REVIEW

  122. Panušková K, Voděrová L, Vaculín Š (2024): Methylphenidate attenuates signs of evoked neuropathic pain in animal model. Physiol Res. 2023 Dec 29;72(S5):S551-S558. doi: 10.33549/physiolres.935215. PMID: 38165759; PMCID: PMC10861255.

  123. http://www.adhs-anderswelt.de/viewtopic.php?p=659654

  124. Martinez, Pasquereau, Drui, Saga, Météreau, Tremblay (2020): Ventral striatum supports Methylphenidate therapeutic effects on impulsive choices expressed in temporal discounting task. Sci Rep. 2020 Jan 20;10(1):716. doi: 10.1038/s41598-020-57595-6. PMID: 31959838.

  125. Steckel (2006): Persönlichkeitsmerkmale erwachsener ADHS-Patienten – Veränderungen unter der Behandlung mit Methylphenidat; Dissertation; n = 24

  126. Shih, Shang, Gau (2018): Comparative Efficacy of Methylphenidate and Atomoxetine on Emotional and Behavioral Problems in Youths with Attention-Deficit/Hyperactivity Disorder. J Child Adolesc Psychopharmacol. 2018 Nov 17. doi: 10.1089/cap.2018.0076.

  127. Blader, Pliszka, Kafantaris, Foley, Carlson, Crowell, Bailey, Sauder, Daviss, Sinha, Matthews, Margulies (2020): Stepped Treatment for Attention-Deficit/Hyperactivity Disorder and Aggressive Behavior: A Randomized, Controlled Trial of Adjunctive Risperidone, Divalproex Sodium, or Placebo After Stimulant Medication Optimization. J Am Acad Child Adolesc Psychiatry. 2020 Jan 30:S0890-8567(20)30064-2. doi: 10.1016/j.jaac.2019.12.009. PMID: 32007604.

  128. Furukawa, Molina da Costa, Bado, Hoefle, Vigne, Monteiro, Wickens, Moll, Tripp, Mattos (2019): Methylphenidate modifies reward cue responses in adults with ADHD: An fMRI study. Neuropharmacology. 2019 Nov 2:107833. doi: 10.1016/j.neuropharm.2019.107833.

  129. Tucha O, Prell S, Mecklinger L, Bormann-Kischkel C, Kübber S, Linder M, Walitza S, Lange KW (2006): Effects of methylphenidate on multiple components of attention in children with attention deficit hyperactivity disorder. Psychopharmacology (Berl). 2006 Apr;185(3):315-26. doi: 10.1007/s00213-006-0318-2. PMID: 16521033.

  130. Jaeschke, Sujkowska, Sowa-Kućma (2021): Methylphenidate for attention-deficit/hyperactivity disorder in adults: a narrative review. Psychopharmacology (Berl). 2021 Oct;238(10):2667-2691. doi: 10.1007/s00213-021-05946-0. PMID: 34436651; PMCID: PMC8455398.

  131. Oades RD, Sadile AG, Sagvolden T, Viggiano D, Zuddas A, Devoto P, Aase H, Johansen EB, Ruocco LA, Russell VA (2005): The control of responsiveness in ADHD by catecholamines: evidence for dopaminergic, noradrenergic and interactive roles. Dev Sci. 2005 Mar;8(2):122-31. doi: 10.1111/j.1467-7687.2005.00399.x. PMID: 15720370. REVIEW

  132. Trott, Wirth (2000): die Pharmakotherapie der hyperkinetischen Störungen; in: Steinhausen (Herausgeber) hyperkinetischen Störungen bei Kindern, Jugendlichen und Erwachsenen, 2. Aufl., Seite 211

  133. Rajala, Populin, Jenison (2020): Methylphenidate affects task-switching and neural signaling in non-human primates. Psychopharmacology (Berl). 2020 May;237(5):1533-1543. doi: 10.1007/s00213-020-05478-z. PMID: 32067136; PMCID: PMC7196524.

  134. Trott, Wirth (2000): die Pharmakotherapie der hyperkinetischen Störungen; in: Steinhausen (Herausgeber) hyperkinetischen Störungen bei Kindern, Jugendlichen und Erwachsenen, 2. Aufl., Seite 212, mit weiteren Nachweisen

  135. Levi-Sachar, Gvirts, Goldwin, Bloch, Shamay-Tsoory, Zagoory-Sharon, Feldman, Maoz (2019): The effect of methylphenidate on social cognition and oxytocin in children with attention deficit hyperactivity disorder. Neuropsychopharmacology. 2019 Sep 12. doi: 10.1038/s41386-019-0522-5.

  136. Rezende, Pacheco, Branco, Fernandes, Boldrini, Doria Filho, Bazán, Amaro Junior, Reed, Casella (2020): Combining neuropsychological tests to improve the assessment of arithmetic difficulties in children with ADHD. Arq Neuropsiquiatr. 2020 Apr;78(4):193-198. doi: 10.1590/0004-282X20190178. PMID: 32130296. n = 42

  137. Golubchik, Manor, Shoval, Weizman (2020): Levels of Proneness to Boredom in Children with Attention-Deficit/Hyperactivity Disorder On and Off Methylphenidate Treatment. J Child Adolesc Psychopharmacol. 2020 Feb 7:10.1089/cap.2019.0151. doi: 10.1089/cap.2019.0151. PMID: 32031873. n = 30

  138. Low, le Sommer, Vangkilde, Fagerlund, Glenthøj, Sonuga-Barke, Habekost, Jepsen (2018): Delay aversion and Executive Functioning in adults with Attention-Deficit/Hyperactivity Disorder: Before and after stimulant treatment. Int J Neuropsychopharmacol. 2018 Aug 16. doi: 10.1093/ijnp/pyy070.

  139. Golubchik, Weizman (2019): Poor performance of the ‘child Reading the Mind in the Eyes Test’ correlates with poorer social-emotional functioning in children with attention-deficit/hyperactivity disorder. Int Clin Psychopharmacol. 2019 Nov 22. doi: 10.1097/YIC.0000000000000299.

  140. Rothe J, Kattlun FA, Kaufmann J, Uhlmann A, Wanderer S, Bluschke A, Beste C, Roessner V (2023): Effects of methylphenidate and physiotherapeutic treatment on graphomotor movements in children with ADHD. Eur Child Adolesc Psychiatry. 2023 Jan 23. doi: 10.1007/s00787-023-02144-5. PMID: 36688969.

  141. Solanto (2001): Dopamine dysfunction in AD/HD: integrating clinical and basic neuroscience research. Behavioural Brain Research 130 (2002) 65–71

  142. Luo, Covey, Hu, Winhusen, Nunes (2019): Differential posttreatment outcomes of methylphenidate for smoking cessation for individuals With ADHD. Am J Addict. 2019 Sep 19. doi: 10.1111/ajad.12961.

  143. Baas, Boot, van Gaal, de Dreu, Cools (2019): Methylphenidate does not affect convergent and divergent creative processes in healthy adults. Neuroimage. 2019 Oct 17:116279. doi: 10.1016/j.neuroimage.2019.116279.

  144. Ten, Tseng, Chiang, Wu, Chen (2019): Creativity in children with ADHD: Effects of medication and comparisons with normal peers. Psychiatry Res. 2019 Nov 5:112680. doi: 10.1016/j.psychres.2019.112680. n = 86

  145. Arnold: Journal of Attention Disorders Vol. 3(4):200-211 (2000) Methylphenidate vs. amphetamine: Comparative review, n = 174 REVIEW

  146. Roskell, Setyawan, Zimovetz, Hodgkins (2014): Systematic evidence synthesis of treatments for ADHD in children and adolescents: indirect treatment comparisons of lisdexamfetamine with methylphenidate and atomoxetine. Curr Med Res Opin. 2014 Aug;30(8):1673-85. doi: 10.1185/03007995.2014.904772.

  147. Prasad, Steer (2008): Switching from neurostimulant therapy to atomoxetine in children and adolescents with attention-deficit hyperactivity disorder : clinical approaches and review of current available evidence. Paediatr Drugs. 2008;10(1):39-47. doi: 10.2165/00148581-200810010-00005. PMID: 18162007. REVIEW

  148. Nagy, Häge, Coghill, Caballero, Adey, Anderson, Sikirica, Cardo (2015): Functional outcomes from a head-to-head, randomized, double-blind trial of lisdexamfetamine dimesylate and atomoxetine in children and adolescents with attention-deficit/hyperactivity disorder and an inadequate response to methylphenidate.Eur Child Adolesc Psychiatry. 2016 Feb;25(2):141-9. doi: 10.1007/s00787-015-0718-0.

  149. Vallejo-Valdivielso, de Castro-Manglano, Díez-Suárez, Marín-Méndez, Soutullo (2020): Clinical and Neuropsychological Predictors of Methylphenidate Response in Children and Adolescents with ADHD: A Naturalistic Follow-up Study in a Spanish Sample. Clin Pract Epidemiol Ment Health. 2019 Dec 31;15:160-171. doi: 10.2174/1745017901915010160. PMID: 32174998; PMCID: PMC7040471.

  150. Michelini, Lenartowicz, Vera, Bilder, McGough, McCracken, Loo SK (2022): Electrophysiological and Clinical Predictors of Methylphenidate, Guanfacine, and Combined Treatment Outcomes in Children With Attention-Deficit/Hyperactivity Disorder. J Am Acad Child Adolesc Psychiatry. 2022 Aug 8:S0890-8567(22)01229-1. doi: 10.1016/j.jaac.2022.08.001. PMID: 35963559. n = 181

  151. Barkley, DuPaul, McMurray (1991): Attention deficit disorder with and without hyperactivity: Clinical response to three dose levels of methylphenidate. Pediatrics, 87, 519–531

  152. Barkley (2001): The inattentive type of ADHD as a distinct disorder: What remains to be done. Clinical Psychology: Science and Practice, 8, 489– 493

  153. Milich, Balentine, Lynam (2001): ADHD combined type and ADHD predominantly inattentive type are distinct and unrelated disorders. Clinical Psychology: Science and Practice, 8, 463– 488

  154. Weiss, Worling, Wasdell: (2003). A chart review study of the inattentive and combined types of ADHD. Journal of Attention Disorders, 7, 1–9 REVIEW

  155. Diamond: Attention-deficit disorder (attention-deficit/hyperactivity disorder without hyperactivity): A neurobiologically and behaviorally distinct disorder from attention-deficit (with hyperactivity), Development and Psychopathology 17 (2005), 807–825, Seite 811

  156. Safren, Perlman, Sprich et al (2008): Kognitive Verhaltenstherapie der ADHS des Erwachsenenalters, Seite 9

  157. Diamond: Attention-deficit disorder (attention-deficit/hyperactivity disorder without hyperactivity): A neurobiologically and behaviorally distinct disorder from attention-deficit (with hyperactivity), Development and Psychopathology 17 (2005), 807–825, Seite 811

  158. Menneh (2008): Prediction of the clinical response to psychostimulant by the basal and reactive salivary cortisol in children with ADHD.

  159. Froehlich, Becker, Nick, Brinkman, Stein, Peugh, Epstein (2018): Sluggish Cognitive Tempo as a Possible Predictor of Methylphenidate Response in Children With ADHD: A Randomized Controlled Trial. J Clin Psychiatry. 2018 Feb 27;79(2). pii: 17m11553. doi: 10.4088/JCP.17m11553.

  160. Tarrant, Roy, Deb, Odedra, Retzer, Roy (2018): The effectiveness of methylphenidate in the management of Attention Deficit Hyperactivity Disorder (ADHD) in people with intellectual disabilities: A systematic review. Res Dev Disabil. 2018 Sep 25;83:217-232. doi: 10.1016/j.ridd.2018.08.017. REVIEW

  161. Sun, Tseng, Wu, Li, Chen, Stubbs, Carvalho, Chen, Lin, Cheng, Wu (2019): Therapeutic effects of methylphenidate for attention-deficit/hyperactivity disorder in children with borderline intellectual functioning or intellectual disability: A systematic review and meta-analysis. Sci Rep. 2019 Nov 4;9(1):15908. doi: 10.1038/s41598-019-52205-6. n = 423 REVIEW

  162. Ogrim, Kropotov, Brunner, Candrian, Sandvik, Hestad (2014): Predicting the clinical outcome of stimulant medication in pediatric attention-deficit/hyperactivity disorder: data from quantitative electroencephalography, event-related potentials, and a go/no-go test. Neuropsychiatr Dis Treat. 2014 Feb 3;10:231-42. doi: 10.2147/NDT.S56600. N = 188

  163. Strauß, Reif, Ulke, Paucke, Sander, Hegerl, Weber, Heupel, Kopf, Kittel-Schneider (2019): Is brain arousal regulation a predictor of response to psychostimulant therapy in adult ADHD patients? Eur Arch Psychiatry Clin Neurosci. 2019 Nov 26. doi: 10.1007/s00406-019-01085-y.

  164. Arnett, Rutter, Stein (2022): Neural Markers of Methylphenidate Response in Children With Attention Deficit Hyperactivity Disorder. Front Behav Neurosci. 2022 May 6;16:887622. doi: 10.3389/fnbeh.2022.887622. PMID: 35600991; PMCID: PMC9121006.

  165. Milich, Balentine, Lynam (2001): ADHD combined type and ADHD predominantly inattentive type are distinct and unrelated disorders. Clinical Psychology: Science and Practice, 8, 463– 488, zitiert nach Diamond: Attention-deficit disorder (attention-deficit/hyperactivity disorder without hyperactivity): A neurobiologically and behaviorally distinct disorder from attention-deficit (with hyperactivity), Development and Psychopathology 17 (2005), 807–825, Seite 811

  166. Traicu, Grizenko, Fortier, Fageera, Sengupta, Joober (2019): Acute blood pressure change with methylphenidate is associated with improvement in attention performance in children with ADHD. Prog Neuropsychopharmacol Biol Psychiatry. 2019 Aug 12;96:109732. doi: 10.1016/j.pnpbp.2019.109732.

  167. Gokten, Tulay, Beser, Yuksel, Arikan, Tarhan, Metin (2019): Predictive Value of Slow and Fast EEG Oscillations for Methylphenidate Response in ADHD. Clin EEG Neurosci. 2019 Sep;50(5):332-338. doi: 10.1177/1550059419863206.

  168. Zhong, Yang, Su, Qian, Cao, Chang, Wang, Yang (2020): The Association with Quantitative Response to Attention-Deficit/Hyperactivity Disorder Medication of the Previously Identified Neurodevelopmental Network Genes. J Child Adolesc Psychopharmacol. 2020 Jul;30(6):348-354. doi: 10.1089/cap.2018.0164. PMID: 32175767. n = 241

  169. Yuan, Zhang, Huang, Wang, Jiao, Huang (2021): Noradrenergic genes polymorphisms and response to methylphenidate in children with ADHD: A systematic review and meta-analysis. Medicine (Baltimore). 2021 Nov 19;100(46):e27858. doi: 10.1097/MD.0000000000027858. PMID: 34797323; PMCID: PMC8601359. n = 1.382

  170. Cascone AD, Calabro F, Foran W, Larsen B, Nugiel T, Parr AC, Tervo-Clemmens B, Luna B, Cohen JR (2023): Brain tissue iron neurophysiology and its relationship with the cognitive effects of dopaminergic modulation in children with and without ADHD. Dev Cogn Neurosci. 2023 Jul 7;63:101274. doi: 10.1016/j.dcn.2023.101274. PMID: 37453207; PMCID: PMC10372187.

  171. Blum NJ, Shults J, Barbaresi W, Bax A, Cacia J, Deavenport-Saman A, Friedman S, Loe IM, Mittal S, Vanderbilt D, LaRosa A, Harstad E. Do Externalizing and Internalizing Symptoms Moderate Medication Response in Preschool Attention-Deficit/Hyperactivity Disorder? A DBPNet Study. J Dev Behav Pediatr. 2023 Sep 1;44(7):e447-e454. doi: 10.1097/DBP.0000000000001209. PMID: 37696030

  172. Cheon KA, Cho DY, Koo MS, Song DH, Namkoong K (2009): Association between homozygosity of a G allele of the alpha-2a-adrenergic receptor gene and methylphenidate response in Korean children and adolescents with attention-deficit/hyperactivity disorder. Biol Psychiatry. 2009 Apr 1;65(7):564-70. doi: 10.1016/j.biopsych.2008.12.003. PMID: 19150055.

  173. Polanczyk, Zeni, Genro, Guimarães, Roman, Hutz, Rohde (2007): Association of the adrenergic alpha2A receptor gene with methylphenidate improvement of inattentive symptoms in children and adolescents with attention-deficit/hyperactivity disorder. Arch Gen Psychiatry. 2007 Feb;64(2):218-24. doi: 10.1001/archpsyc.64.2.218. PMID: 17283289.

  174. Cho SC, Kim BN, Cummins TD, Kim JW, Bellgrove MA. (2012): Norepinephrine transporter -3081(A/T) and alpha-2A-adrenergic receptor MspI polymorphisms are associated with cardiovascular side effects of OROS-methylphenidate treatment. J Psychopharmacol. 2012 Mar;26(3):380-9. doi: 10.1177/0269881111405356. PMID: 21628343. n = 101

  175. Song J, Song DH, Jhung K, Cheon KA (2011): Norepinephrine transporter gene (SLC6A2) is involved with methylphenidate response in Korean children with attention deficit hyperactivity disorder. Int Clin Psychopharmacol. 2011 Mar;26(2):107-13. doi: 10.1097/YIC.0b013e32834152d1. PMID: 21127421. n = 114

  176. Yang L, Wang YF, Li J, Faraone SV (2004): Association of norepinephrine transporter gene with methylphenidate response. J Am Acad Child Adolesc Psychiatry. 2004 Sep;43(9):1154-8. doi: 10.1097/01.chi.0000131134.63368.46. PMID: 15322419.

  177. Kim BN, Kim JW, Hong SB, Cho SC, Shin MS, Yoo HJ. (2010): Possible association of norepinephrine transporter -3081(A/T) polymorphism with methylphenidate response in attention deficit hyperactivity disorder. Behav Brain Funct. 2010 Oct 7;6:57. doi: 10.1186/1744-9081-6-57. PMID: 20929549; PMCID: PMC2959002. n = 112

  178. Myer NM, Boland JR, Faraone SV (2018): Pharmacogenetics predictors of methylphenidate efficacy in childhood ADHD. Mol Psychiatry. 2018 Sep;23(9):1929-1936. doi: 10.1038/mp.2017.234. PMID: 29230023; PMCID: PMC7039663. METASTUDY

  179. Shi J, Xiao J, Wang X, Jung SM, Bleske BE, Markowitz JS, Patrick KS, Zhu HJ. (2022): Plasma Carboxylesterase 1 Predicts Methylphenidate Exposure: A Proof-of-Concept Study Using Plasma Protein Biomarker for Hepatic Drug Metabolism. Clin Pharmacol Ther. 2022 Apr;111(4):878-885. doi: 10.1002/cpt.2486. PMID: 34743324; PMCID: PMC9249567.

  180. https://www.adhs365.de/adhs/medikamente/kartoffelm%C3%BCsli/

  181. https://adhs-forum.adxs.org/t/wirkstoff-probleme/6920/16; Forumsanmeldung erforderlich